You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/235259061

Comparison of ceramic and polymeric membrane permeability and fouling


using surface water

Article  in  Separation and Purification Technology · June 2011


DOI: 10.1016/j.seppur.2011.03.025

CITATIONS READS

125 3,243

5 authors, including:

Bas Hofs Julien Ogier


Evides Lanxess AG
26 PUBLICATIONS   685 CITATIONS    9 PUBLICATIONS   150 CITATIONS   

SEE PROFILE SEE PROFILE

Dirk Vries Erwin Beerendonk


KWR Water Research Institute KWR Water Research Institute
35 PUBLICATIONS   292 CITATIONS    42 PUBLICATIONS   1,688 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Optimal (experimental) input design View project

Control of infinite dimensional systems View project

All content following this page was uploaded by Bas Hofs on 07 December 2017.

The user has requested enhancement of the downloaded file.


Separation and Purification Technology 79 (2011) 365–374

Contents lists available at ScienceDirect

Separation and Purification Technology


journal homepage: www.elsevier.com/locate/seppur

Comparison of ceramic and polymeric membrane permeability and fouling using


surface water
Bas Hofs ∗ , Julien Ogier, Dirk Vries, Erwin F. Beerendonk, Emile R. Cornelissen
KWR Watercycle Research Institute, P.O. Box 1072, 3430 BB Nieuwegein, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: The trans membrane pressure (TMP) increase at a constant flux of 150 L m−2 h−1 due to membrane fouling
Received 11 January 2011 by direct filtration with lake water is investigated for four ceramic (Al2 O3 , ZrO2 , TiO2 , SiC) and a PES/PVP
Received in revised form 2 March 2011 polymeric microfiltration membrane(s). The membrane structures and compositions are characterised
Accepted 2 March 2011
with permporometry (pore size, porosity) and XPS. The TiO2 and SiC membrane have a 5 and 24 times
larger average pore size than expected based on supplier information.
Keywords:
The TMP increase rate due to fouling inversely follows the measured pore size. Reversible fouling
Microfiltration
decreases in the order of polymeric ≈ Al2 O3 ≈ ZrO2 > TiO2 > SiC, and for irreversible fouling poly-
Natural organic matter
Low pressure membranes
meric > ZrO2 > Al2 O3 > TiO2 > SiC.
Ultrafiltration Removal of non purgeable organic carbon (NPOC) is around 30% for the ceramic membranes, and
Water treatment 13–25% for the polymeric membrane. The reversible NPOC load decreases in the following order
(Al2 O3 > (ZrO2 ≈ TiO2 )) ≈ SiC > polymeric. The higher degree of fouling on the polymeric membrane is
partly due to its lower volume/area ratio, compared to ceramic membranes. Mass balance analysis shows
that 85 ± 8% of the NPOC is accounted for. Thus, the used soaking method (pH 12 NaOH, >1 h) does not
fully remove the NPOC, suggesting that a more aggressive cleaning solution should be used.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction ceramic membranes is evident from applied high backwash pres-


sures [11–13], although ceramic membranes are more prone to
Ceramic membranes have been hailed for their advantageous breakage than polymeric membranes. The chemical stability of
properties when compared to polymeric membranes. The advan- ceramic membranes is higher than that of polymeric membranes,
tages of ceramic membranes compared to polymeric membranes and decreases in the order of TiO2 , ZrO2 , Al2 O3 , SiO2 and their
are often stated to be: (i) a relatively narrow pore size distribution hydrothermal stability decreases in the order of ZrO2 , Al2 O3 , TiO2 ,
and higher porosity, resulting in better separation characteristics SiO2 [14,15]. The stability also depends on the membrane struc-
and a higher flux, (ii) a higher mechanical stability (allowing higher ture, with finer structures being less stable. The hydrophilicity of
pressures), (iii) a higher chemical stability resulting in longer mem- polymeric membranes depends on the used polymers and can be
brane lifetimes, and (iv) higher hydrophilicity resulting in high high; some polymeric membranes have a contact angle with water
fluxes at low pressures [1–7]. Higher hydrophilicity is also asso- of 10◦ [16]. For ceramic membranes the hydrophilicity depends
ciated with lower fouling [6]. Evidence to support all of these on the used materials. The contact angle of water for ceramic
advantages of ceramic membranes is, however, not so easily found. membranes covers a broad range, and the same can be said of poly-
A study in which the pore size distributions (usually measured by meric membranes [9,10,17], making it hard to determine whether
fractional rejection of molecules) of a thin film composite and a ceramic membranes are indeed more hydrophilic than polymeric
TiO2 membrane are compared, shows that they are very similar ones.
[8]. Unfortunately, no further studies in which pore size distribu- Differences in fouling behavior of polymeric and ceramic mem-
tions are measured for both polymeric and ceramic membranes branes can be expected, as the surface groups of ceramic and
within a single article could be found. The pore size distribution polymeric membranes are different. In the scientific literature,
of ceramic membranes can be narrow [9], but the same is true studies comparing the fouling of ceramic and polymeric mem-
for polymeric membranes [10]. The higher mechanical stability of branes for direct water filtration could not be found. However, some
work has been done within the Techneau framework [18]. Mueller
et al. [19,20] compared the fouling of a ceramic membrane (TiO2 top
layer, Al2 O3 support) with that of a polymeric (polyethersulfone)
∗ Corresponding author. Tel.: +31 0 306069697; fax: +31 0 306061165. membrane, both operated under the same conditions, by analysing
E-mail address: bas.hofs@kwrwater.nl (B. Hofs). the cleaning in place (CIP, at pH 11) waste from natural organic

1383-5866/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.seppur.2011.03.025
366 B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374

matter (NOM) loaded membranes by liquid chromatography tubular membrane for the fouling experiments) were obtained
organic carbon detection (LC-OCD) and fluorescence excitation- from Cometas (Bagsvaerd, Denmark). The polymeric filter pen
emission matrices (FEEM). LC-OCD showed that the CIP-waste of membrane, composed of a PES/PVP blend, was obtained from Norit
the polymeric membrane showed a higher polysaccharide content Filtrix (Amersfoort, The Netherlands). Pore sizes and the materials
than the ceramic membrane [20]. FEEM showed that the CIP-waste as provided by the suppliers (for the top layer) of the membranes
of the polymeric membrane had higher fluorescence for protein- are given in Tables 1 and 2. Note that due to the varying require-
like organic compounds, and the CIP-waste from the ceramic ments of the characterisation devices we used different membrane
membrane showed higher fluorescence for humic substances [19]. dimensions for each test. The pore size of the membranes as given
As polysaccharides are thought to be a major foulant for microfiltra- by the supplier was similar for the individual membranes. Details
tion/ultrafiltration (MF/UF) [21,22], ceramic membranes may have about the membranes used in the filtration test are given in Table 1,
a lower potential for organic fouling [20]. However, more research and for characterisation by porometry, XPS and zeta-potential mea-
into fouling mechanisms of polymeric and ceramic membranes surements are given in Table 2.
is needed to prove whether polysaccharides lead to irreversible The pore size and pore size distribution of the ceramic mem-
fouling. branes was measured with a Porolux 1000 capillary flow porometer
In general, pretreatment is often employed when low pressure (IB-FT GmbH, Wildau, Germany). The membranes were cut into
membranes are selected. Studies show that 30–40% of the low pres- samples with a length of approximately 7 cm and were fitted on one
sure membranes operate without pretreatment, and that especially end with a PVC collar of 12 mm outer diameter and a Swagelok fer-
coagulation and powdered activated carbon are commonly applied rule set. The other end was closed by potting with a two component
as pretreatment for MF/UF [23,24]. Coagulation is the most success- resin. Porofil (Benelux Scientific, Tiel, The Netherlands) wetting
ful pretreatment for fouling control, but may increase fouling and fluid was used to wet the membrane prior to the measurements,
thus requires an optimized dosing of the used chemicals [24]. For which were carried out in duplicate. The mean pore diameter is
these reasons we decided to study the performance of the mem- calculated from the mean flow pressure, which corresponds to the
branes without coagulation, or any other form of pretreatment – intersection of the wet curve with the calculated half-dry curve
although in a future study we intend to incorporate this aspect as (calculated from a dry and a wet run).
well. The porosity was measured on an Accupyc 1330 pycnometer
In this study, we investigate the performance of four different from Micromeritics instruments Corporation (Norcross, U.S.A.). For
ceramic membranes (TiO2 , ZrO2 , Al2 O3 , and SiC) and one poly- these measurements the membranes were cut into small pieces
meric membrane (polyethersulfone–polyvinylpyridine, PES–PVP). of about 7 mm length. A sample was sealed in a sample compart-
The performance and fouling of the membranes by surface water ment of known volume. Helium was added, and then expanded
are compared under the same flux and backwash procedure. In into another calibrated volume. The pressure before and after the
practice, in full scale application, the membranes would be used cell expansion was measured to compute the sample volume. The
under different operational conditions than here. The ceramic porosity was subsequently calculated from the density.
membranes are usually operated at a higher flux than the poly- The elemental composition of the membranes was measured on
meric, and the backwash and cleaning procedures would also be a Quantera SXM (scanning XPS microprobe) from Physical Electron-
very different for the ceramic and polymeric membranes. However, ics (Chanhassen, U.S.A.). The XPS measurements were done with an
since we want to compare the membrane properties on the level Al K␣ monochromatic (at 1486.6 eV) X-ray beam, with a beam size
of the membranes rather than different operational conditions, we of 100 ␮m (power 25 W, Ie 2.6 mA). The Quantera SXM was con-
decided to fix the flux and backwash procedure. trolled with Compass software and the data reduction was done
Thus, this paper is on the comparison of the membrane proper- with Multipak v8.0. Samples of the membranes of approximately
ties with respect to trans membrane pressure increase (TMP) and 7 mm length were used for the XPS measurements. Elemental spec-
characterisation of fouling by surface water. Therefore we charac- trum scans of the both top and support layer (3 locations) were
terise the membranes in order to verify the membrane properties measured (at two points per location), averaged and interpreted
as supplied by the suppliers. We characterise the water type and with the help of standard books.
follow the TMP due to both reversible and irreversible fouling in fil- All the above measurements were carried out at the Euro-
tration experiments at a constant flux. The reversible fouling is the pean Membrane Institute (EMI) at the University of Twente (The
TMP increase within one filtration cycle, and the irreversible fouling Netherlands).
is determined from the increase in the TMP directly after the back- Zeta-potential measurements were performed using the zeta
wash. The following membrane properties were obtained: pore meter SurPass (Anton Paar GmbH, Austria). These measurements
size distribution, porosity, zeta-potential, and elemental composi- were performed on a flat sheet membrane of SiC and on seven chan-
tion of top and support layer. The composition (Ca2+ concentration, nels tubular membranes of Al2 O3 , TiO2 and ZrO2 . A special adapter
conductivity, LC-OCD, turbidity, etc.) of the feed water type was was used to force the measuring electrolyte solutions (pH 6–10,
measured, as well as the non-purgeable organic compound (NPOC) 10−3 M KCl) through the porous wall that separate two individ-
content of the backwash water, permeate and the CIP-waste. This ual channels. The zeta potential was calculated from the streaming
allows us to make a mass balance of the NPOC and to determine potential measurements as described elsewhere [25].
the separation characteristics of the membranes.
2.2. Water type and characterisation

2. Materials and methods The first set of membrane filtration tests at fixed backwash
pressure was performed with surface water from Noord-Bergum
2.1. Membranes and characterisation lake (The Netherlands). A batch of 400 L was sampled (on the 20th
of November 2009 and subsequently homogenised and filtered
The Al2 O3 , TiO2 and ZrO2 single channel tubular membranes through a 1 ␮m cartridge filter. The water was stored at 4 ◦ C and
were obtained from Atech innovations GmbH (Gladbeck, Germany) the water composition is given in Table 3 (1st batch). The amount
and had a support layer of Al2 O3 . The SiC membranes (a flat disc of NPOC found is nearly the same as the amount of DOC, due to
type for the pore size, porosity, X-ray photoelectron spectroscopy pre-filtering though a 1 ␮m cartridge filter. The second set of exper-
(XPS) and zeta-potential measurements, and a single channel iments at comparable backwash flux was performed with surface
B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374 367

Table 1
Membranes that were used for the filtration tests.

Type Compositiona Supplier Pore sizea Surface areaa Configuration

ZrO2 L = 250 mm
Al2 O3 Atech Innovations 0.1 ␮m 0.0047 m2 Dint = 6 mm
Ceramic
TiO2

SiC Cometas 0.04 ␮m 0.013 m2 L = 305 mm


Dint = 14 mm

Polymeric PES/PVP Norit Filtrix 0.14 ␮m 0.02 m2 L = 200 mm


Dint = 0.4 mm
a
Supplier’s information, L = length, Dint = internal diameter. Note that suppliers may use a different method to determine the pore size of their membranes than the method
used here.

Table 2
Membranes that were used for the characterisation (porometry, XPS, and zeta-potential).

Compositiona Membranes used for pore size, porosity and XPS measurements Membranes used for zeta-potential measurements

ZrO2 L = 250 mm L = 200 mm


Dint = 6 mm Dint = 6 mm
Al2 O3 Dext = 10 mm, Dext = 25.4 mm,
TiO2 1 channel 7 channels

SiC Flat membrane Flat membrane


D = 25 mm 55 mm × 25 mm
Thickness = 6 mm Thickness = 6 mm

Polymeric 1 channel, L = 200 mm –


Dint = 0.72 mm
Dext = 1.2 mm
a
Supplier’s information, L = length, Dint = internal diameter, Dext = external diameter.

water from Noord-Bergum lake (The Netherlands) sampled on the inductively coupled plasma mass spectrometry (ICP-MS) at the
2nd of September 2010, homogenised and filtered through a 1 ␮m KWR laboratory (Nieuwegein, The Netherlands) following an in-
cartridge filter. The water was stored at 4 ◦ C and the water compo- house procedure (LAM-058). All other (Table 3) measurements
sition is given in Table 3 (2nd batch). were performed at the KWR laboratory, with in-house methods or
Liquid Chromatography-Organic Carbon Detection (LC-OCD) text-book procedures (turbidity: in accordance with NEN-ISO 7027,
was performed by the DOC-Labor (Germany). Sulphate was deter- UV absorbance: at 254 nm, in-house method LAM-033, hydrogen
mined by Vitens Laboratory Utrecht (Utrecht, The Netherlands). carbonate: by titration with acid, in-house method LAM-042, elec-
NPOC was determined in accordance with ISO 8245. Calcium, trical conductivity: measured with a CDM 83 conductivity meter
magnesium and chloride concentrations were determined with (Radiometer), conforming to ISO 7888, at 25 ◦ C).

Table 3
Feed water composition, after 1 ␮m cartridge filter, as determined before the experiments.

1st batch 2nd batch Unit

pH 7.9 7.7 –
Conductivity 690 640 ␮S/cm
Calcium 49 42 mg/L
Magnesium 14 12
Chloride 110 110
Iron 0.59 0.28
Hydrogen carbonate 125 115
Sulphate 55 55
Turbidity 9.7 3.6 FNU
NPOC 15 14 mg/L
UV 41 32 E/m
DOCa 15.1 – mg/L
HOCa nq –
POCa nq –
CDOCa 15.1 –
Humic acidsa 10.3 –
Building blocksa 2.0 –
Acidsa nq –
Neutralsa 2.0 –
Biopolymers 0.76 –
SAC 0.01 – m−1
SUVA 3.14 – L mg−1 m−1

HOC = hydrophobic organic carbon, CDOC = chromatographic DOC.nq = not quantifiable (<1 ppb). NPOC and turbidity of the first batch of water were determined after the
first set of experiments as well; NPOC was the same, but turbidity was slightly lower (9.2 FNU).
a
Data from LC-OCD measurement.
368 B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374

permeate given by (Vb × cb ) + 1/Nb ((Vs × cs ) + (Vp × cp )) = 1/Nb (Vf × cf ), where


cooler
Nb is the number of backwash cycles in one experiment, V is the
pressured air volume, and c is the concentration of the NPOC of the stream
Membrane
module
denoted by the subscripts – b, s, p, and f, which represent back-
Feed Backwash wash, soaking solution, permeate and feed, respectively. The
tank tank concentration of NPOC and turbidity and volume of the back-
wash was corrected for the amount of feed water and permeate
P
4-way that inevitably was mixed with the backwash water (backwash
valve
with permeate). The percentage found by the mass balance is
((Vb × cb ) + 1/Nb ((Vs × cs ) + (Vp × cp )0/(1/Nb (Vf × cf )) × 100. A similar
pump mass balance for particles cannot be made based on the turbidity
measurements performed in this study.
Sewage

Fig. 1. Scheme of lab-scale pilot in filtration mode. In backwash mode the 4-way 2.3.2. Test at comparable backwash flux
valve is rotated around its central point by 90◦ . Additional tests with the TiO2 and the polymeric membrane
were done to elucidate the effect of the variation in the backwash
flux. The tests were performed as described above with the fol-
2.3. Experimental filtration setup and procedure lowing differences. A new batch of Noord Bergum lake water was
obtained (2nd batch, Table 3), and a new membrane was used for
Two series of experiments were performed. In the first set each test. A number of the channels in the polymeric membrane
the backwash pressure was kept constant as described below. In were blocked with two-component resin, decreasing the surface
the second set the backwash pressure or surface area was mod- area of the membrane to 0.0045 m2 . The feed flow was lowered to
ified in order to get comparable backwash fluxes, as described 11.3 mL/min in order to maintain the same flux in all experiments.
below. The duration of the filtration tests with the polymeric membrane
was about 5 h. For the test with the TiO2 membrane, the backwash
2.3.1. Tests at constant backwash pressure pressure was lowered by decreasing the air pressure. The cumu-
Tests were performed with a lab-scale pilot, see Fig. 1. A TSE lative backwash production and cumulative permeate production
540270 dual syringes pump (TSE systems GmbH, Bad homburg, were stored and analysed for the additional tests with the TiO2
Germany) provided an accurate constant feed flow during all the and polymeric membrane. For the calculation of the mass balances
experiments. The surface areas of the membranes were not iden- Nb = 1 (see the formulas in Section 2.3.1).
tical, so the feed flow was adjusted to 11.8, 11.8, 11.8, 33.5, and
50 mL/min, for the ZrO2 , Al2 O3 , TiO2 , SiC and polymeric membrane, 2.3.3. Data handling
respectively, in order to reach a constant feed flux of 150 L m−2 h−1 . Reversible and irreversible fouling was determined from oper-
A pressure sensor in the feed of the set-up (Fig. 1) measured the ational data. For the reversible fouling the decrease in the TMP by
pressure in front of the membrane, which equals TMP which was a backwash was used, averaged over the total number of backwash
recorded on a computer every 10 s. A vessel containing fresh MilliQ cycles. The TMP at time t (TMPt ) minus the TMP at t = 0 (TMP0 ) was
water, pressurised at 3 bars was used for the backwash of mem- plotted versus the permeate volume to determine the irreversible
branes. The 1 min backwash occurred automatically every 20 min. fouling, which was obtained by the slope of a linear fit to the TMPs
All the experiments were carried out in duplicate and started just after each of the backwashes (see Fig. 2B).
by MilliQ filtration (150 L m−2 h−1 , with backwash as described) to
remove air from the system, until a constant TMP was reached to 2.3.4. Membrane cleaning procedure
obtain a reference TMP value for each test. Then the water was After a filtration test, the ceramic membrane was cleaned by
switched from MilliQ to Noord-Bergum surface water. The subse- soaking; first in citric acid (1%) for a minimum of 2 h and then
quent filtration test lasted about 5 h. In the tests with the polymeric in sodium hypochlorite (3000 ppm) for a minimum of 2 h. After-
membrane the duration of the filtration was shorter (2.5 h) and wards, the membrane was rinsed with MilliQ water and stored
could not be continued due to the fact that the maximum pump in MilliQ water until the next experiment. Before starting a new
pressure (about 1.3 bar) was reached. All tests were performed at test the ceramic membrane was characterised with MilliQ water
20 ◦ C and the temperature was controlled by a Julabo FC1200 cooler at 150 L m−2 h−1 to verify TMP0 . A new measurement was only
system (JULABO Labortechnik GmbH, Seelbach, Germany). started if the pressure had not changed. The polymeric membrane
The last backwash sample collected after 5 h of filtration was was not cleaned; a new module was used for each test. The addi-
stored, as well as the permeate sample of the 20 min filtration tional tests at different backwash pressures were performed with
cycle before this backwash. Directly after the final backwash, the new membranes.
membrane was removed from its housing and soaked in sodium
hydroxide solution (pH 12) for at least 1 h. Those three samples, 3. Results and discussion
i.e. last backwash water after 5 h, permeate before this backwash
and soaking solutions, were weighed and NPOC, UV and turbid- 3.1. Membrane characterisation
ity were measured. The NPOC mass balance was based on these
samples. The polymeric membrane cannot withstand pH 12, so no The used ceramic membranes were characterised by
soaking sample was obtained for the polymeric membrane exper- permporometry, XPS and zeta-potential measurements (Table 4).
iment. We assume that the NPOC collected from the backwash For the polymeric membrane the measured pore size is the same
is part of the reversible fouling, whilst the NPOC collected from as that given by the supplier. For the Al2 O3 , and ZrO2 ceramic
soaking at pH 12 is part of the irreversible fouling. The concentra- membranes the measured mean flow pore size deviates a factor
tion of NPOC, the turbidity and the UV values of the feed water, 2.4 and 1.6, respectively, from the pore size in the supplier infor-
backwash water, permeate and CIP waste were measured, as well mation. For TiO2 and especially SiC the deviation is larger; a factor
as the volume (by measuring the mass) of the collected samples. of 5 and 24, respectively. The SiC membrane is very hydrophilic
The NPOC mass balance is based on a single backwash and thus (resulting in a water up-take when in contact with water) which
B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374 369

A 0.8
may have influenced the used method, as in permporometry a
liquid is displaced from the membrane. However, in the literature
no references were found reporting on shifts to larger pore size as
0.6
determined with porometry in connection with hydrophilicity of
TMP (bar)

the membrane.
0.4 The porosity of the membranes is comparable, 72–78%, except
for the SiC membrane which has a much lower porosity of 40%
0.2 (Table 4). A clean water flux obtained with milliQ at 150 L m−2 h−1
resulted in an initial TMP, of 0.042, 0.073, 0.10, 0.14, and 0.19 bar,
0.0 for the TiO2 , SiC, polymeric, Al2 O3 , and ZrO2 membrane, respec-
0 5000 10000 15000 20000 tively. Initially, we also selected a SiO2 membrane for this study.
Time (s) During the clean water flux tests however, a steady decrease in
B TMP was observed for the SiO2 membrane, probably as a result of
0.8
a slow dissolution of the membrane layer of the SiO2 membrane
[26]. Therefore the investigation with this ceramic membrane
TMPt - TMP0 (bar)

0.6 was aborted. This problem was not encountered with the other
membranes.
0.4 For the SiC membrane the iso-electric point was determined
from zeta-potential measurements and found to be in accor-
0.2 dance with the literature values (Table 4). Unfortunately, repeated
y = 1.60*10-4x + 0.071 attempts to determine the iso-electric point of the other ceramic
2
R = 0.965 membranes were unsuccessful. We assume that the values found
0.0
in literature for particles and/or surfaces made of the same material
0 200 400 600 800
Produced permeate (L/m2) as the ceramic membranes are representative.
The membrane composition was determined with XPS analysis,
Fig. 2. (A) Typical fouling experiment, from which reversible fouling was deter- to be able to compare the investigated membrane composition to
mined. (B) Determination of irreversible fouling for the Al2 O3 membrane. The literature data and supplier information. Furthermore, the elemen-
straight line in Fig. 2B is a linear fit to the first point of each filtration cycle (after
tal composition of the membranes facilitates analysis of variations
the backwash), omitting the starting point of the first filtration cycle. Flux was
150 L m−2 h−1 . in different membrane batches. The presence of the expected
elements was confirmed in the different layers, but there were
significant deviations from expectations based on supplier infor-
mation (Table 5). For the ZrO2 and Al2 O3 membranes there were
some small deviations. The presence of C for instance, is probably

Table 4
Membrane properties – comparison of measurements and information from supplier and literature.

Type Pore size (␮m) Iso-electric point Volume


porositya (%)

Measureda Supplier Measured Literature


j
Al2 O3 0.24 0.1 8–9.4b , c , d 73
j
TiO2 0.51 0.1 5.1–6.4b , c 72
j
ZrO2 0.16 0.1 6.3–7.1b , e , f 74
SiC 0.96 0.04 2.6 2.5–3.5d , g , h 40
Polymeric 0.14 0.14 – 4–5i 78
a
Mean flow pore size and porosity were determined by porometry as described in the Section 2 (Section 2.1).
b
[27].
c
[28].
d
[29].
e
[30].
f
[31].
g
[32].
h
[33].
i
According to supplier.
j
Could not be determined.

Table 5
Membrane composition in % as determined by XPS.

Location of analysis Analysis C O Al Si Ti Zr

ZrO2 Top layer 13.9 55 9 16.7


Support layer 15.6 53.2 21.4 3.6 1.7 0.2
TiO2 Top layer 20.9 51.1 7.4 0.9 11.5
Support layer 14.4 52.7 22.7 4 2
Al2 O3 Top layer 8.8 58.1 26.1 2.7 0.12
Support layer 10.5 54.1 26.2 4.4 2
SiC Top layer 24.2 45.6 3.5 26.7
Support layer 20.0 53.0 2.1 24.9

The expected composition follows from the formula given in de first column (e.g. for ZrO2 , Zr is 33 and O is 66). Some other elements (Na, Mg, Ca) were also found, but in
low amounts (generally <2%, at maximum about 5%).
370 B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374

due to organic contaminations from air [34]. The found percentages Table 6
Volume/Area ratio of the membranes and used backwash fluxes.
(9–24%) are in agreement with other XPS measurements on ceramic
membranes [35]. For the ZrO2 top layer we find 16.7% Zr and 9% Membrane V/A (m) Backwash flux (×103 L m−2 h−1 )
Al. This gives an expected O-content of 16.7 × 2 + 9 × 1.5 = 46.9%; Al2 O3 1.9 1.2
whereas 55% was measured. The relatively high value for O in the TiO2 1.9 1.5 or 2.8
top layer can be explained by the presence of contaminants as indi- ZrO2 1.9 0.57
cated by the amount of C found; it is likely that the contaminants SiC 3.6 1.0
Polymer 0.28 0.33 or 1.3
consist partly of C and O. Another contribution to a higher O-content
than expected is the fact that the surface of ZrO2 has a higher O-
content than the bulk. The XPS shows that support layer of the
ZrO2 membrane is made up predominantly of O and Al, in accor- had a large variation in the backwash flux however, it may have
dance with supplier information. However, contaminations (high played a significant role in influencing the fouling, especially for
C-content, higher O-content than expected, and some Si) are also the polymeric and TiO2 membrane as the backwash fluxes were
present on/in the support layer. Also, it is clear that the support relatively small and large (0.33 and 2.8 × 103 L m−2 h−1 ), respec-
layer is already visible in the analysis of the top layer. tively. Therefore we performed another series of experiments with
The top layers of the TiO2 and Al2 O3 membrane also show a these membranes with more comparable backwash fluxes, nar-
relatively high O-content. For the TiO2 and Al2 O3 membranes we rowing the range of backwash flux to 4–10 times (Table 6). It is
would expect 36.0 and 44.6% O, respectively – from the presence of found that lowering the backwash flux for the TiO2 membrane
Al, Ti and Si – where 51 and 58% are found, respectively. The support from about 2800–1500 L m−2 h−1 decreases reversible fouling from
layers also show a higher O-content than expected. Again, these 1.2 ± 0.01 to 0.08 ± 0.02 bar – note that this is barely significant –
higher values are probably due to the presence of contaminants on and does not affect irreversible fouling significantly (4.3 ± 0.5 vs.
the surface and support layer of the membrane. The high level of O- 4.3 ± 1.4 × 10−5 bar m2 L−1 ). For the polymeric membrane, increas-
content found for the SiC membrane can again be due to the above ing the backwash flux from about 330–1300 L m−2 h−1 has no
mentioned contamination. Alternatively, the membrane could have significant effect on reversible fouling, but seems to decrease irre-
adsorbed water molecules, as the SiC membrane is very hydrophilic versible fouling. Furthermore the filtration run lasted about 5 h with
(a dry membrane takes up water when put into contact with it), or the higher backwash flux whereas with the lower flux it had to be
some SiO2 -groups. stopped after 2.5 h due to reaching the maximum reachable pres-
sure (pump). However, the observed differences with variation in
3.2. Fouling experiments backwash flux are small compared to the differences between the
different types of ceramic and polymeric membranes. Apart from
3.2.1. TMP increase this, the minor differences could be caused by the lower turbidity
An increase in the feed pressure due to reversible and irre- of the second batch of water compared to the first batch.
versible fouling of an Al2 O3 membrane due to direct filtration with Regardless of the backwash flux, the reversible fouling depends
the Noord-Bergum lake water was observed (Fig. 2A). Reversible on membrane type and follows the following series: poly-
fouling was determined from Fig. 2A, by calculating the average meric ≈ Al2 O3 ≈ ZrO2 > TiO2 > SiC (Fig. 3). The irreversible fouling
increase in TMP within all filtration cycles. The irreversible fouling also depends on the membrane type and follows the following
in bar m2 L−1 was determined from the slope of the fouling curve series: polymeric > ZrO2 > Al2 O3 > TiO2 > SiC. There seems to be a
as shown graphically in Fig. 2B. Reproducibility of the fouling mea- relation between the measured pore size of the membranes and
surements with the different membranes, expressed as reversible the increase in TMP due to reversible and irreversible fouling. The
and irreversible fouling and based on a duplicate measurement was membranes with the highest measured pore size show the low-
usually within 10%, except for the ZrO2 membrane. We were unable est increase in TMP due to fouling. The SiC membrane, which has
to clean the ZrO2 membrane completely with the described clean- the lowest iso-electric point and highest hydrophilicity and the
ing protocol. The initial TMP of the ZrO2 membrane with milliQ was highest measured pore size, has a TMP increase due to irreversible
0.19 bar with the virgin membrane, this increased to 0.25 bar after fouling of only (2 ± 8) × 10−6 bar m2 L−1 (average ± standard devia-
the first fouling test and cleaning. The reversible and irreversible tion). This could be seen as a confirmation of the hypothesis that the
fouling of the ZrO2 membrane increased in further tests; we only membrane with the highest negative charge or hydrophilicity fouls
show results for the first test in this paper. the least [10,37,38], even though other studies show that this is a
The fouling experiments were performed with the same water
type for all the experiments. The TMP0 was 0.042, 0.082, 0.073,
0.003
0.10, 0.085, 0.14, and 0.19 bar, for the TiO2 , TiO2 *, SiC, polymeric, reversible
Reversible fouling (bar)

0.4
polymeric*, Al2 O3 , and ZrO2 membrane, respectively. The linear fits
Irreversible fouling

irreversible
to determine the irreversible fouling had a R2 of 0.95 ± 0.03 (aver-
(bar*m /L)

0.3 0.002
age ± standard deviation) for all membranes, except the SiC. For
2

the SiC membrane the TMP increase due to irreversible fouling was
limited, resulting in a low R2 (about 0.3). 0.2
Initially experiments were performed with a fixed back- 0.001
wash pressure, but additional experiments were done to obtain 0.1
a set of experiments with comparable backwash flux (about
600–1500 L m−2 h−1 , see Table 6). Criteria for an efficient backwash 0 0
er
C
3

*
2*

procedure include; sufficiently high pressure (about 3–4 times fil-


er
O

Si

m
O

m
Zr
Ti
2

ly
Ti
Al

ly

tration pressure) [36] and sufficient volume to carry the fouling


Po

Po

from the membrane to the backwash vessel. Usually a backwash Membrane type
flux of about 2–3 times the filtration flux is used. We meet the
Fig. 3. TMP increase due to reversible and irreversible fouling. TiO2 *, Poly-
criteria for efficient backwash as the backwash flux used varied
mer* = results with these membranes of tests performed with second batch of water
from 2 to 20 times the filtration flux and the backwash volume was at comparable; backwash flux (see Table 6). ZrO2 results are based on the first
at least 4 times the volume of the tubing + membrane. Since we non-reproducible test with the virgin membrane.
B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374 371

simplification [39,40] and that other factors like surface rough- 100
ness and membrane structure (average pore size and porosity) and
geometry play an important role as well [41–43].
75
The irreversible fouling found in this work is comparable to val-

Removal (%)
ues in literature; irreversible fouling of a ceramic membrane by NPOC
a synthetic water after coagulation was found to vary from about 50 Turbidity
10−5 to about 10−6 bar m2 L−1 [44], and irreversible fouling of a UV
PES/PVP UF membrane by surface water was found to be about 25
10−4 bar m2 L−1 [45]. These two papers [44,45] use different meth-
ods to calculate the irreversible fouling. We basically followed the
method by Meyn et al. [44]. Recalculating the irreversible fouling 0

er
C
3

*
2*

er
O

O
obtained in this study to conform to the method used by Abrahamse

Si

ym
O

m
Zr
Ti
2

Ti
Al

ly
l
et al. [45], gives values ranging from 2 × 10−6 to 2 × 10−4 bar m2 L−1

Po

Po
(for SiC to polymeric, respectively). The latter method thus gives Membrane type
smaller values for irreversible fouling than the method employed Fig. 4. Removal of NPOC, turbidity and UV by the different membranes. TiO2 *, Poly-
in this study. mer* = results with these membranes of tests performed with second batch of water.
Another factor that plays a role in fouling of membranes is the Shown are average and standard deviation (error bar). Where no error bar is shown,
geometry of the membranes [43]. The volume over area ratio, V/A the duplicate measurement resulted in the same values (except for ZrO2 , where no
duplicate measurement could be performed).
(Table 6), for the Al2 O3 , TiO2 and ZrO2 membranes is the same,
1.9, since the membranes have the same geometry. When V/A is
relatively high, the concentration of accumulated foulants during a
using porometry is not well suited for this membrane. This is in
filtration cycle in the membrane (and possibly tubing in front of the
contradiction with the results obtained for the TMP increase due to
membrane) remains relatively low. Fouling might than be expected
fouling, which correlate rather well with the measured pore size.
to be lower for membranes with a higher V/A and vice versa. The
In principle, the rejection of negatively charged NPOC may be
polymeric membrane has the lowest V/A, 0.28, and should therefore
increased by the presence of negative charges on the membrane.
be expected to have the highest fouling, which is indeed observed
The SiC membrane is negatively charged as it has an iso-electric
(Fig. 3). The V/A of the polymeric membrane is a factor 7 lower than
point of 2.6. Fouling probably decreases the charge on the SiC mem-
that of the Al2 O3 , TiO2 and ZrO2 membranes, whilst the difference
brane, but it probably remains negatively charged. Based on data
in fouling is a factor of 1–4 for reversible fouling and 10–50 for
in Table 3 and assuming the missing cations necessary for elec-
irreversible fouling. We hypothesize that the differences found here
troneutrality are Na+ , we estimate the ionic strength of the water is
in fouling between the polymeric and Al2 O3 or ZrO2 membranes
about 9 mM. This gives a Debye screening length [49] (the distance
can mostly be attributed to the differences in V/A.
over which the electrostatic potential drops to 1/e of the origi-
The SiC membrane has the highest V/A of the used membranes,
nal value) of about 3 nm. If the pore size of the SiC membrane is
3.6, and shows the lowest fouling (Fig. 3). The difference between
larger than that of the other membranes, for example 0.2 ␮m, the
V/A for the SiC and the Al2 O3 , TiO2 and ZrO2 membranes is a factor of
pore diameter would be only slightly decreased by electrostatic
2, whilst the difference in fouling is a factor of 10–30 for reversible
screening (about 10%; in the case described above, 10 nm from the
fouling and 20–100 for irreversible fouling. It seems unlikely that
edge of the pore the electrostatic potential has dropped to about
the difference V/A alone can explain the observed difference in
5% of its value at the edge). The electrostatic effect therefore is not
fouling for the SiC membrane.
expected to decrease the effective pore size significantly in these
experiments and is expected to play only a minor role in removal
3.2.2. Removal efficiency
of the negatively charged fraction of NOM for this membrane.
Removal of particles based on turbidity is more than 95% for
Besides the properties of the membrane, also the properties
all membranes except for the SiC membrane, which shows 92%
of the fouling layer are expected to influence the removal effi-
removal. This suggests that the pore size of the SiC membrane is
ciencies (see Section 3.2.3). Since here we analysed the removal
larger than the pore size of the other membranes, as shown by the
efficiency at the end of the tests, the charge on the membranes is
measurement of the pore size (Section 3.1).
influenced by the fouling. We did not measure to which extent the
Removal of NPOC is around 30% for the ceramic membranes, and
fouling changes the charge size and/or sign. If one wants to esti-
13–25% for the polymeric membrane (Fig. 4). The removal of 13%
mate the effect of charge on the removal efficiency of NPOC, these
for a polymeric membrane is comparable with results published for
data (obtainable by e.g. streaming potential measurements on the
membranes with similar pore sizes [46–48], and so is the removal of
membrane) would be needed.
30% for the ceramic membranes [2,11]. The low removal of 13% was
found for the 2.5 h long experiment with the polymeric membrane,
whereas a higher value of 25% was found when the test lasted the 3.2.3. Fouling load
full 5 h (which was the duration of the test with the ceramic mem- The NPOC load of the membranes was measured for reversible
branes). Apparently the removal increases with the duration of the fouling by collecting the backwash water of the last filtration cycle,
test, indicating that the increase of accumulated irreversible fouling and for irreversible fouling by analysing CIP soaking water (see Sec-
increases the NPOC removal efficiency. UV removal shows simi- tion 2.3). For the ZrO2 membrane we were unable to measure a
lar trends as the NPOC removal. The lower UV and NPOC removal duplicate and only the results of the first test are given as an indi-
for the polymeric membrane compared to the ceramic membranes cation. The reversible fouling load decreases in the following order;
could on one hand be expected, as the polymeric membrane has a (Al2 O3 > ZrO2 ≈ TiO2 ) ≈ SiC > polymeric (Fig. 5).
slightly larger pore size compared to ceramic membranes, accord- The irreversible fouling load after 5 h of operation seems highest
ing to the suppliers. On the other hand, it is unexpected as the pore for Al2 O3 and ZrO2 and lowest for TiO2 and SiC, but the error bars
size measurements give a very large pore size for the SiC membrane for the latter two are rather large. Differences between the NPOC
compared to the value given by the supplier (0.9 ␮m and 0.04 ␮m, load of the different membranes are within a factor of 1.5, whereas
respectively). As the SiC membrane shows the same removal as the the differences in the TMP increase due to fouling are much more
other membranes, this indicates that the pore size determination pronounced (Fig. 3).
372 B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374

0.10 0.30 0.25 0.30


reversible load

Irreversible load (g/m )


Reversible load (g/m )

Irreversible load (FNU/m )


Reversible load (FNU/m )
2
2

2
2
0.08 irreversible load
0.20
0.20
0.06 0.20
0.15
0.04
0.10 0.10
0.02 0.10

0.05
0.00 0.00 reversible load
irreversible load

er
C
3

*
2*

er
O

Si 0.00 0.00

m
O

m
Ti

Zr
2

ly
Ti
Al

ly
Po

Po

er
C
3

*
2*

er
O

Si

m
O

m
Zr
Ti
2

ly
Membrane type

Ti
Al

ly
Po

Po
Membrane type
Fig. 5. NPOC load of the membranes. Note that the NPOC load of the polymeric
membrane is after 2.5 h of filtration, whereas for the other membranes (including
Fig. 6. Turbidity load of the membranes. Note that the turbidity load of the poly-
polymer*) it is after 5 h. TiO2 *, Polymer* = results with these membranes of tests
meric membrane is after 2.5 h of filtration, whereas for the other membranes
performed with second batch of water.
(including polymer*) it is after 5 h. TiO2 *, Polymer* = results with these membranes
of tests performed with second batch of water.

Mass balance analysis (see Section 2.3) shows that 85 ± 8% (aver-


age + standard deviation) of the total NPOC in the feed water is advanced method to determine the particle load, which can distin-
accounted for in the backwash water, soaking solution and perme- guish the particles’ sizes and particle size distribution is needed
ate. Mass balance analysis shows that 88, 87, 90, 68, and 82% of the to assess the composition of the particles in the reversible and
expected NPOC was accounted for, for the Al2 O3 , TiO2 , TiO2 *, ZrO2 , irreversible fouling. With this information, a TMP increase may
and SiC membrane, respectively. This shows that a significant part possibly be related to a particle load parameter and membrane
of the fouling NPOC is probably still attached to the membranes type.
after soaking in 0.01 M NaOH. Especially the ZrO2 membrane is not For a thorough comparison of membrane fouling using different
cleaned well. Additionally, for the ZrO2 membrane the CIP removed types of membranes, the following membrane properties should
only a part of the irreversible NPOC load, which is shown by the preferably be measured or known: membrane chemistry, mem-
observed non-reproducibility of the fouling experiment. The low brane geometry, membrane permeability or resistance, pore size
mass balance percentage of NPOC accounted for can also be due to distribution, porosity, contact angle, zeta-potential of the support
the fact that we based our analysis on the last filtration cycle. It is and the top layer, thickness of the top layer, and surface roughness.
likely that NPOC removal increased with increasing filtration time The composition (concentrations of various ions, ionic strength,
due to the build up of a fouling layer that acts as an extra barrier. DOC, LC-OCD, turbidity, zeta-potential) of the feed water type
However, for the experiments with comparable backwash flux we should also be measured, as well as the composition of the back-
collected all the backwash water and the results of the mass bal- wash water and the fouling layer. We did not determine the surface
ance were similar. Thus, it seems that not all NPOC was removed roughness, the thickness of the top layer and the contact angles of
from the membranes by soaking in a NaOH solution (pH 12), even the membranes, nor did we perform an extensive analysis of the
though this method is used for cleaning of membranes [50]. This reversible and irreversible fouling composition. In order to under-
suggests that for ceramic membranes, a more extensive or aggres- stand the differences in fouling a more detailed analysis is needed.
sive treatment is necessary to remove all the NPOC, for example a However, this is beyond the scope of the investigation described
more concentrated NaOH solution (e.g. pH 13) at 50 ◦ C [51]. here.
No clear relation seems to exist between the NPOC load of foul-
ing and the TMP increase due to reversible or irreversible fouling,
although the membranes with the lowest irreversible NPOC load 4. Conclusions
(the SiC and TiO2 membranes), are also the ones that show the
least irreversible fouling characteristics (compare Figs. 3 and 5). We compared the permeability and fouling of four different
The fact that no clear relation seems to exist between either NPOC ceramic membranes and one polymeric membrane using surface
load or turbidity load and fouling shows that the combined effect water.
of NOM and particles determine the TMP increase due to fouling Pore size measurements by permporometry showed that the
[52]. pore size of the Al2 O3 and ZrO2 were similar to the values given
The turbidity load of the membranes is shown in Fig. 6. by the supplier. The measured pore size for the TiO2 and SiC mem-
The reversible turbidity load is highest for the TiO2 membrane brane deviated by a factor of 5 and 24 from those given by the
(although the error bar shows this might not be very signifi- supplier, respectively. The properties of the membranes may cause
cant), which is advantageous since this turbidity fraction is easily deviations in the used method; the SiC membrane is exceptionally
removed with a backwash and does not contribute to long term hydrophilic.
irreversible fouling. One might expect a higher reversible turbidity XPS measurements confirmed the membrane compositions as
load for the SiC membrane, but this membrane did not remove as documented by the suppliers, although the measurements showed
much turbidity as the other membranes (see Fig. 4). The level of that all membrane samples were contaminated with carbon and
reversible turbidity load is apparently sufficiently high to prevent oxygen.
any pressure build-up due to irreversible fouling. Furthermore, less The TiO2 and especially the SiC membrane showed a low TMP
turbidity will build up because of the potentially more open struc- increase due to low reversible and irreversible fouling, compared
ture of the SiC membrane as shown by the pore size measurements. to the Al2 O3 , ZrO2 and polymeric membranes. For the polymeric
Unfortunately, a relation between the turbidity load and the membrane the differences can probably be attributed to differences
TMP increase could not be identified. In this respect it is noted in the V/A ratio between the membrane modules. The TMP increase
that turbidity is a very indirect measure of particle load. A more due to fouling increases with decreasing measured pore size.
B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374 373

Removal of NOM and UV was lowest for the polymeric mem- [15] T. Van Gestel, C. Vandecasteele, A. Buekenhoudt, C. Dotremont, J. Luyten,
brane (13–25%) and higher and comparable (around 30%) for the B. Van der Bruggen, G. Maes, Corrosion properties of alumina and titania
NF membranes, Journal of Membrane Science 214 (2003), PII S0376-
different ceramic membranes. The removal of turbidity was >95% 7388(0302)00513-00517.
for all membranes except the SiC membrane. [16] N. Park, B. Kwon, I.S. Kim, J. Cho, Biofouling potential of various NF membranes
The reversible NPOC fouling load decreases in the following with respect to bacteria and their soluble microbial products (SMP): character-
izations, flux decline, and transport parameters, Journal of Membrane Science
order: (Al2 O3 > (ZrO2 ≈ TiO2 )) ≈ SiC > polymeric and the irreversible 258 (2005) 43–54.
fouling load after 5 h of operation decreases in the order; [17] A. Dobrak, B. Verrecht, H. Van den Dungen, A. Buekenhoudt, I.F.J. Vankelecom,
Al2 O3 ≈ ZrO2 > TiO2 ≈ SiC. The load of reversible and irreversible B. Van der Bruggen, Solvent flux behavior and rejection characteris-
tics of hydrophilic and hydrophobic mesoporous and microporous TiO2
NPOC fouling, however, was not very different for the Al2 O3 , TiO2 ,
and ZrO2 membranes, Journal of Membrane Science 346 (2010) 344–
ZrO2 , and SiC membranes, suggesting that these parameters cannot 352.
be used as an indicator of TMP increase. [18] http://www.techneau.org.
[19] U. Mueller, R. Schaefer, C. Zawadsky, K. Boeckle, M. Boesl, I. Eberhagen,
Mass balance analysis of the NPOC load on the membranes, in
B. Hambsch, Techneau Report D 2.3.2.7. Removal of phages and nanopar-
the permeate and in the backwash, showed that 85 ± 8% of the ticles by ceramic membranes, in: DVGW Water technology Centre (TZW),
expected NPOC could be accounted for. Probably, the soaking at 2009.
pH 12 did not remove a significant amount of the fouling materials [20] U. Mueller, R. Schaefer, M. Witte, K. Boeckle, M. Boesl, I. Eberhagen, B. Hambsch,
Techneau Report D 2.3.2.2, in: Interim Report Removal of particulate mat-
from the membranes. This is especially true for the ZrO2 mem- ter by ceramic membranes during surface water treatment in, DVGW Water
brane where only 68% of the NPOC was accounted for. This suggests technology Centre (TZW), 2008.
that a more aggressive CIP is necessary for the complete removal [21] L. Fan, J. Harris, F. Roddick, N. Booker, Fouling of microfiltration membranes by
the fractional components of natural organic matter in surface water, Water
of irreversible fouling. Science and Technology: Water Supply (2002) 313–320.
[22] N. Lee, G. Amy, J.P. Croue, H. Buisson, Identification and understanding of fouling
in low-pressure membrane (MF/UF) filtration by natural organic matter (NOM),
Acknowledgements Water Research 38 (2004) 4511–4523.
[23] J.M. Laine, D. Vial, P. Moulart, Status after 10 years of operation – overview of
This research was conducted in the framework of the Techneau UF technology today, Desalination 131 (2000) 17–25.
[24] H. Huang, K. Schwab, J.G. Jacangelo, Pretreatment for low pressure membranes
consortium (financed by the EU) and the Joint Research Program in water treatment: a review, Environmental Science and Technology 43 (2009)
(BTO) of the Dutch drinking water companies. We thank one 3011–3019.
reviewer for his/her comments, which helped to greatly improve [25] M. Wegmann, B. Michen, T. Luxbacher, J. Fritsch, T. Graule, Modifi-
cation of ceramic microfilters with colloidal zirconia to promote the
the quality of this paper. We thank Z. Borneman and H.A. Teunis adsorption of viruses from water, Water Research 42 (2008) 1726–
(both from the EMI, TU Twente, The Netherlands) and T. Luxbacher 1734.
(from Anton Paar, Graz, Austria) for their efforts to characterise the [26] D. Wolff-Boenisch, S.R. Gislason, E.H. Oelkers, C.V. Putnis, The dissolution rates
of natural glasses as a function of their composition at pH 4 and 10.6 and
membranes.
temperatures from 25 to 74 ◦ C, Geochimica Et Cosmochimica Acta 68 (2004)
4843–4858.
[27] M. Kosmulski, Compilation of PZC and IEP of sparingly soluble metal oxides
References and hydroxides from literature, Advances in Colloid and Interface Science 152
(2009) 14–25.
[1] H.K. Oh, S. Takizawa, S. Ohgaki, H. Katayama, K. Oguma, M.J. Yu, Removal of [28] M. Mullet, P. Fievet, J.C. Reggiani, J. Pagetti, Surface electrochemical properties
organics and viruses using hybrid ceramic MF system without draining PAC, of mixed oxide ceramic membranes: zeta-potential and surface charge density,
Desalination 202 (2007) 191–198. Journal of Membrane Science 123 (1997) 255–265.
[2] C. Moulin, M.M. Bourbigot, A. Tazipain, F. Bourdon, Design and performance of [29] X.H. Wang, Y. Hirata, Colloidal processing and mechanical properties of SiC with
membrane filtration installations – capacity and product quality for drinking- Al2 O3 and Y2 O3 , Nippon Seramikkusu Kyokai Gakujutsu Ronbunshi/Journal of
water applications, Environmental Technology 12 (1991) 841–858. the Ceramic Society of Japan 112 (2004) 22–28.
[3] T. Tsuru, Inorganic porous membranes for liquid phase separation, Separation [30] Y.K. Leong, N. Katiforis, D.B.O. Harding, T.W. Healy, D.V. Boger, Role of rheol-
and Purification Methods 30 (2001) 191–220. ogy in colloidal processing of ZrO sub 2, Journal of Materials Processing and
[4] R.J. Ciora Jr., P.K.T. Liu, Ceramic membranes for environmental related applica- Manufacturing Science 1 (1993) 445–453.
tions, Fluid: Particle Separation Journal 15 (2003) 51–60. [31] M. Mao, D. Fornasiero, J. Ralston, R.S.C. Smart, S. Sobieraj, Electrochemistry
[5] B. Hofs, S.G.J. Heijman, J.Z. Hamad, M. Kennedy, G. Amy, Ceramic microfiltra- of the zircon–water interface, Colloids and Surfaces A: Physicochemical and
tion with a sub-micron PAC pre-coat for water treatment, in: J.P.T. Melin, M. Engineering Aspects 85 (1994) 37–49.
Dohmann (Eds.), Aachener Tagung Wasser und Membranen, vol. 8, Aachen, [32] S. Sano, T. Banno, M. Maeda, K. Oda, Y. Shibasaki, Slip casting and sintering of
2009, p. W12. silicon carbide (part 1) – slip preparation of silicon carbide powder produced by
[6] B. Van Der Bruggen, C. Vandecasteele, T. Van Gestel, W. Doyen, R. Leysen, A Acheson method, Nippon Seramikkusu Kyokai Gakujutsu Ronbunshi/Journal of
review of pressure-driven membrane processes in wastewater treatment and the Ceramic Society of Japan 104 (1996) 984–988.
drinking water production, Environmental Progress 22 (2003) 46–56. [33] S.H. Yeh, C.C. Wan, Codeposition of SiC powders with nickel in a Watts bath,
[7] E.R. Cornelissen, B. Hofs, U. Muller, E.F. Beerendonk, S.G.J. Heijman, Direct and Journal of Applied Electrochemistry 24 (1994) 993–1000.
hybrid ceramic microfiltration in water treatment, in: TECHNEAU: Safe Drink- [34] D.R. Cousens, B.J. Wood, J.Q. Wang, A. Atrens, Implications of specimen prepara-
ing Water from Source to Tap, IWA Publishing, Maastricht, 2009, 83–97. tion and of surface contamination for the measurement of the grain boundary
[8] S. Lee, G. Park, G. Amy, S.K. Hong, S.H. Moon, D.H. Lee, ChoF J., Determination of carbon concentration of steels using X-ray microanalysis in an UHV FESTEM,
membrane pore size distribution using the fractional rejection of nonionic and Surface and Interface Analysis 29 (2000) 23–32.
charged macromolecules, Journal of Membrane Science 201 (2002) 191–201. [35] M.I. Vazquez, R. de Lara, J. Benavente, Characterisation of two microporous
[9] Q. Zhang, Y. Fan, N. Xu, Effect of the surface properties on filtration performance ceramic membranes with different geometrical parameters: effect of protein
of Al2 O3 –TiO2 composite membrane, Separation and Purification Technology adsorption on electrochemical parameters, Journal of the European Ceramic
66 (2009) 306–312. Society 27 (2007) 4245–4250.
[10] G.N.B. Barona, B.J. Cha, B. Jung, Negatively charged poly(vinylidene fluoride) [36] M. Kennedy, S.M. Kim, I. Mutenyo, L. Broens, J. Schippers, Intermittent
microfiltration membranes by sulfonation, Journal of Membrane Science 290 cross flushing of hollow fiber ultrafiltration systems, Desalination 118 (1998)
(2007) 46–54. 175–188.
[11] A. Lerch, S. Panglisch, P. Buchta, Y. Tomitac, H. Yonekawa, K. Hattori, R. Gimbel, [37] B. Kaeselev, J. Pieracci, G. Belfort, Photoinduced grafting of ultrafiltration
Direct river water treatment using coagulation/ceramic membrane microfil- membranes: comparison of poly(ether sulfone) and poly(sulfone), Journal of
tration, Desalination 179 (2005) 41–50. Membrane Science 194 (2001) 245–261.
[12] T. Meyn, A. Bahn, T.O. Leiknes, Significance of flocculation for NOM removal [38] H. Huang, T. Young, J.G. Jacangelo, Novel approach for the analysis of bench-
by coagulation–ceramic microfiltration, Water Science and Technology: Water scale, low pressure membrane fouling in water treatment, Journal of Membrane
Supply 8 (2008) 691–700. Science 334 (2009) 1–8.
[13] N. Shirasaki, T. Matsushita, Y. Matsui, K. Ohno, Effects of reversible and irre- [39] M. Taniguchi, J.E. Kilduff, G. Belfort, Low fouling synthetic membranes
versible membrane fouling on virus removal by a coagulation–microfiltration by UV-assisted graft polymerization: monomer selection to mitigate foul-
system, Journal of Water Supply: Research and Technology: AQUA 57 (2008) ing by natural organic matter, Journal of Membrane Science 222 (2003)
501–506. 59–70.
[14] A. Buekenhoudt, Stability of Porous Ceramic Membranes, in: R. Mallada, M. [40] M. Zhou, H. Liu, J.E. Kilduff, R. Langer, D.G. Anderson, G. Belfort, High-
Menendez (Eds.), Inorganic membranes; synthesis, characterization and appli- throughput membrane surface modification to control NOM fouling,
cations, Elsevier, Amsterdam, 2008, pp. 1–31. Environmental Science and Technology 43 (2009) 3865–3871.
374 B. Hofs et al. / Separation and Purification Technology 79 (2011) 365–374

[41] E.M. Vrijenhoek, S. Hong, M. Elimelech, Influence of membrane surface prop- [46] A.I. Schaefer, A.G. Fane, T.D. Waite, Cost factors and chemical pretreatment
erties on initial rate of colloidal fouling of reverse osmosis and nanofiltration effects in the membrane filtration of waters containing natural organic matter,
membranes, Journal of Membrane Science 188 (2001) 115–128. Water Research 35 (2001) 1509–1517.
[42] C.R. Reiss, J.S. Taylor, C. Robert, Surface water treatment using nanofiltra- [47] M. Siddiqui, G. Amy, J. Ryan, W. Odem, Membranes for the control of nat-
tion – pilot testing results and design considerations, Desalination 125 (1999) ural organic matter from surface waters, Water Research 34 (2000) 3355–
97–112. 3370.
[43] S. Panglisch, Formation and prevention of hardly removable particle layers in [48] N. Lee, G. Amy, J. Lozier, Understanding natural organic matter fouling in low-
inside-out capillary membranes operating in dead-end mode, Water Science pressure membrane filtration, Desalination 178 (2005) 85–93.
and Technology: Water Supply (2003) 117–124. [49] http://en.wikipedia.org/wiki/Debye length.
[44] T. Meyn, T.O. Leiknes, Comparison of optional process configurations and [50] N. Porcelli, S. Judd, Chemical cleaning of potable water membranes: a review,
operating conditions for ceramic membrane MF coupled with coagula- Separation and Purification Technology 71 (2010) 137–143.
tion/flocculation pre-treatment for the removal of NOM in drinking water [51] M. Bartlett, M.R. Bird, J.A. Howell, An experimental study for the development
production, Journal of Water Supply: Research and Technology: AQUA 59 of a qualitative membrane cleaning model, Journal of Membrane Science 105
(2010) 81–91. (1995) 147–157.
[45] A.J. Abrahamse, C. Lipreau, S. Li, S.G.J. Heijman, Removal of divalent cations [52] A.I. Schaefer, U. Schwicker, M.M. Fischer, A.G. Fane, T.D. Waite, Microfiltration
reduces fouling of ultrafiltration membranes, Journal of Membrane Science 323 of colloids and natural organic matter, Journal of Membrane Science 171 (2000)
(2008) 153–158. 151–172.

View publication stats

You might also like