You are on page 1of 13

Progress in Polymer Science 34 (2009) 969–981

Contents lists available at ScienceDirect

Progress in Polymer Science


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

Crosslinked poly(vinyl alcohol) membranes


Brian Bolto ∗ , Thuy Tran 1 , Manh Hoang 1 , Zongli Xie 1
CSIRO Materials Science and Engineering, Private Bag 33, Clayton South MDC, Victoria 3169, Australia

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

Article history: The inherent hydrophilicity of poly(vinyl alcohol) (PVA) makes it an attractive polymer for
Received 13 April 2009 water treatment applications based on membranes. Thermal and chemical resistance and a
Received in revised form 19 May 2009
high anti-fouling potential are accompanied by high water permeability. The large swelling
Accepted 29 May 2009
capacity requires that the PVA be adequately crosslinked to ensure that the contaminants
Available online 7 June 2009
in water can be retained, and that compaction under pressure can be minimised. There is a
challenge to achieve this and still obtain economical permeate fluxes.
Keywords:
Poly(vinyl alcohol) The literature on crosslinking of PVA is reviewed. Many reagents have been explored. Glu-
Crosslinking agents taraldehyde is a more effective crosslinking agent than formaldehyde or glycidyl acrylate,
Reverse osmosis which in turn gives a less swollen product than that obtained by increasing the crystallinity
Pervaporation by heating. Toluene diisocyanate and acrolien give similar results in the preparation of
Ultrafiltration reverse osmosis membranes, but at an extremely high applied pressure. Crosslinking with
Microfiltration maleic anhydride/vinyl methyl ether copolymers gives as good a result, but at even higher
pressure. Thus the high swelling of PVA can be overcome by crosslinking reactions, but
with the consumption of some of the OH groups responsible for the hydrophilicity. What
is really needed is network formation that provides a tight restraining without serious
loss of hydrophilic behaviour. Similar membranes are used for the separation of organic
compounds from one another or from water by pervaporation, where the vapor of one com-
ponent is selectively transferred through the membrane on the basis of polarity differences.
Here PVA membranes would be especially suited to dehydration procedures.
The high swelling behaviour can be countered also by forming the active PVA component
inside the pores of a microporous membrane. Crosslinked PVA inside such membranes
has its swelling suppressed, and can function as salt removal membranes. By only coating
the pore walls, leaving some porosity, microfiltration or ultrafiltration membranes can be
prepared. In both situations a degree of grafting to the host membrane would be beneficial.

Crown Copyright © 2009 Published by Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
2. Crosslinking methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
2.1. Freeze–thaw inducement of crystallisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
2.2. Heat treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
2.3. Acid-catalysed dehydration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 971
2.4. Irradiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 971

∗ Corresponding author. Tel.: +61 3 9545 2037; fax: +61 3 9545 1128.
E-mail addresses: brian.bolto@csiro.au (B. Bolto), thuy.tran@csiro.au (T. Tran), manh.hoang@csiro.au (M. Hoang), zongli.xie@csiro.au (Z. Xie).
1
Fax: +61 3 9545 1128.

0079-6700/$ – see front matter. Crown Copyright © 2009 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2009.05.003
970 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

2.5. Radical production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 971


2.6. Formaldehyde . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 971
2.7. Glutaraldehyde . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 972
2.8. Other aldehydes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 973
2.9. Di-, tri- and polycarboxylic acids, anhydrides or acid chlorides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 974
2.10. Alkoxysilanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 975
2.11. Other chemical crosslinking agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 976
3. Grafting of PVA onto support membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 976
4. Commercial PVA membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 977
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 978
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 978

1. Introduction 2. Crosslinking methods

Crosslinked poly(vinyl alcohol) or PVA has good chem- 2.1. Freeze–thaw inducement of crystallisation
ical, thermal and mechanical stability. It is appropriate
for pressure-driven membranes designed for a variety of PVA can be physically crosslinked by methods that intro-
water treatment applications, such as microfiltration (MF) duce crystalline regions that act as crosslinks [12]. Cycling
for removing particulate material, microbial cell residues at −20◦ /+25 ◦ C for up to 10 times yields membranes that
and turbidity, ultrafiltration (UF) for taking out large contain from 5.2 to 11.6% of crystalline PVA on a swollen
organic molecules, nanofiltration (NF) for small organic basis. Size exclusion effects are observed with theophylline
molecules removal and softening, and reverse osmosis (RO) (MW 180 Da) and a dextran derivative (MW 4400 Da).
for desalination. The highly polar nature of PVA minimizes Although there was no apparent change in properties after
fouling in such applications, as it is well established that several months, the physical crosslinks were not as strong
non-polar surfaces encourage adsorption of water con- or as stable as chemical ones. The water content has been
taminants because of hydrophobic interactions, whether shown to decrease with increasing crystallinity [13].
natural organic compounds such as humic and fulvic acids
[1] or microbial matter [2]. 2.2. Heat treatment
Because of their oleophobic properties, PVA membranes
are useful in wastewater treatment. They are also used for Fully hydrolysed PVA samples of MW 86 and 115 kDa
product recovery and for the separation of organic com- have been heated over the range of 120–175 ◦ C for
pounds from one another or from water by pervaporation 30–80 min and the products tested as RO membranes [14].
(PV), where the vapor of one component is selectively Water and salt permeabilities decreased abruptly on heat
transferred through the membrane on the basis of polar- treatment in operations at 6.7 MPa, with the salt result
ity, not the volatility difference. PVA membranes would be decreasing greater than that for water, indicating that
especially suited to dehydration procedures. the membrane structure became tighter on more intense
The membranes cover a wide range of porosities and heating. Higher temperatures are known to produce unsat-
permeation properties, reflecting a spectrum of degrees uration, chain scission and chemically bonded crosslinking.
of crosslinking, from the minimum necessary to lessen The water contents of the membranes were not reported.
swelling in the more porous membranes to the highly Earlier work showed that PVA membranes baked at 145 ◦ C
crosslinked PVA necessary for rejection of small inorganic gave 22 and 26% salt rejection at 4 and 5.2 MPa respectively
ions in tighter, non-porous RO membranes. However, the [15].
more swollen membranes can become compacted when Interpolymer anionic composite RO membranes have
high pressures are applied, so that flux through the mem- been made from PVA (MW 75 kDa) and poly(styrene sul-
brane is markedly reduced over time. phonic acid) or PSSA (MW 123 kDa) by heat treating cast
Because of its hydrophilic nature, PVA must be modified membranes at 120 ◦ C for 2 h [16]. The optimum ratio of
to minimise swelling in water when fabricated for aqueous PVA to PSSA was 1.5, when the salt rejection was 93% at an
applications. A burgeoning literature exists on this aspect applied pressure of 8 MPa. Infrared studies showed that a
[3–5]. The stability of crosslinked PVA in highly acidic or dehydration reaction produced –O–SO2 – bridges between
highly alkaline environments has been demonstrated [6]. the molecules. Water fluxes were 9–28 L/m2 h, decreasing
While the literature on applications as membranes or sup- with longer heat treatment times and higher PVA con-
port material in water treatment is spasmodic [7], there tents.
has been a continuing effort in the biotechnology area Heat treatment of PVA (99% hydrolysed and MW 75 kDa)
to use PVA membranes for protein recovery [8–10]. As at temperatures up to 160 ◦ C has shown that hygroscop-
well, PVA gels have been studied extensively as bioma- icity diminishes and the rate of water vapor transmission
terials for artificial kidney and pancreas, glucose sensors, decreases, as does the permeability [17]. This was ascribed
immuno-isolation membranes, artificial cartilage, contact to a change in crystallinity rather than the introduction of
lenses and drug delivery systems [11]. Methods of improv- covalent crosslinks.
ing the mechanical integrity of PVA include freezing, heat To avoid the possible leaching of unreacted crosslinker,
treatment, irradiation, and chemical crosslinking. studies have been made of UF membranes made from PVA
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 971

crosslinked by heat treatment at 100 ◦ C for 1 h [11]. Mem- [20]. The swelling ratio (swollen/dry) of the samples in
brane selectivities and fouling by protein were compared water was lowered from 5.7 to 2.6 by the treatment. As
with polyether sulphone (PES) and regenerated cellulose RO membranes at 6.7 MPa, the water permeability and salt
membranes. The PVA membrane was more resistant to rejection were comparable to those of untreated PVA, but
fouling and was considered a viable replacement for the although the water permeability was considerably higher
other membranes. than obtained with chemically crosslinked PVA, the salt
The flux for pure water passage through porous UF rejection was much lower at 20–50%. The radiation prod-
membranes made from heat treated PVA have been ucts hence had a looser, more open structure. Their water
compared with those prepared with a chemical crosslink- contents were over 10 times that of chemically crosslinked
ing agent in the form of glutaraldehyde [18]. Solutions PVA.
10% in PVA (99% hydrolysed and MW 77 kDa) and 0.5%
in poly(ethylene glycol) were cast into membranes and 2.5. Radical production
immersed immediately into an acetone coagulation bath.
The dried membranes were kept in a 120 ◦ C oven for dif- Persulphate treatment of PVA in the form of K2 S2 O8 has
ferent times, giving products with increasing crystallinity been used to insolubilise the polymer [21]. This strongly
longer heat treatment times: 18.7% after 1 h, 27.9% after 2 h oxidising radical producer promotes coupling between
and 48.9% after 3 h. The porosity decreased from 0.48 to polymer radicals, leading to a crosslinked structure. The NF
0.39. For the membrane of highest crystallinity the water membrane formed had rejections of up to 80% in treating
flux decreased from 0.8 to 0.4 L/m2 h after 3 h operation 1000 mg/L MgSO4 . Manipulating the variables can produce
at 200 kPa, versus 0.7–0.4 L/m2 h for the glutaraldehyde- membranes with RO properties, as when aqueous solu-
crosslinked membrane. This was ascribed to swelling of the tions of PVA and K2 S2 O8 are deposited on PES support
membrane in water, then further deformation because of membranes [22]. Crosslinking was completed by heating
the transmembrane pressure, causing irreversible changes at 60–100 ◦ C. Operation was possible at 2 MPa, when the
to the morphology. The heat-treated membranes were salt rejection was 45% at 42 L/m2 h.
claimed to have better anti-compression properties than
the crosslinked membranes because of the supporting 2.6. Formaldehyde
function of the crystal regions versus the fragile acetal
bonds at the points of crosslinking. In the wet state the PVA (99% hydrolysed) has been used to prepare
tensile strength and strain ratio both increase for the heat- membranes that have been crosslinked with excess
treated membranes, but both decrease for the crosslinked formaldehyde in a bath at room temperature for 20 h, then
membranes. For the former membranes the elasticity was heat treated [23]. The degree of formalisation was esti-
greater also. mated to be not greater than 50% because of the difficulty of
formaldehyde penetrating crystallites formed during dry-
2.3. Acid-catalysed dehydration ing. Heat treatment at temperatures up to 200 ◦ C for 10 min
before the formalising step decreased the water absorption
Coating PES or polytetrafluoroethylene support mem- and water permeability of the membrane. With no prior
branes with a solution of PVA and sulphuric acid, heating heating, formalising at 50 ◦ C for 2 h gave membranes with
to 150 ◦ C, then coating with a solution of poly(acrylic acid) 93–97% salt rejection in RO tests at 4 MPa. Similar experi-
and sulphuric acid and heating to 160 ◦ C, produces a com- ments were carried out later to give a membrane that had
posite membrane that is capable of high salt rejections, but 90% salt rejection from simulated seawater at 10 MPa [24].
relatively low fluxes at a pressure of 0.4 MPa [19], as quoted PVA of MWs 45 and 90 kDa have been made from films
by Immelman et al. [4]. placed in a precipitating bath of various salts for an hour,
A PVA solution (86–89% hydrolysed and MW 72 kDa) then 2–3 days in a crosslinking bath of strongly acidic 30%
containing a dehydration catalyst in the form of con- formaldehyde [25]. In RO tests at 5 MPa, rejections of mono-
centrated sulphuric acid has been deposited on an valent salts were 25–28% at a flux of 35–54 L/m2 h. For
asymmetrical PES support membrane, then heated at divalent cations the figures were 23–27% at 41–69 L/m2 h,
100–125 ◦ C for 10–30 min to effect insolubilisation [4]. and for sulphates 69–70% at 42–70 L/m2 h. The results sug-
Ether crosslinks were formed between the PVA chains, but gest that the membranes are negatively charged. This could
water elimination within the chain could produce olefinic arise from incorporation of sulphonic acid groups from the
double bonds, a reaction favoured by higher temperatures sulphuric acid catalyst. Salt rejections of 85% are reported
and indicated by discolourisation. The products were tested for similar membranes operated at 4 MPa [26].
on 2000 mg/L NaCl at 2 MPa and gave salt rejections of up PVA crosslinked with formaldehyde/HCl is useful in
to 70–85% at a flux of 29–33 L/m2 h for the tightest mem- PV of aqueous acetic acid for dehydration purposes
brane prepared under more forcing conditions, versus 55% [27]. The membranes produced had greater fluxes but
rejected at 63 L/m2 h for one with a more open network that lower selectivities than those formed by crosslinking with
was prepared under milder conditions. glutaraldehyde, which was ascribed to more efficient
crosslinking in the case of glutaraldehyde. The degree of
2.4. Irradiation formaldehyde reaction with PVA (99.5% hydrolysed and
MW 77 kDa) has been studied in regard to PV behaviour in
Cast PVA films have been irradiated by a 60 Co irradiation the dehydration of benzene [28]. It was postulated that the
unit at doses of 0.5–40 Mrad, at a dose rate of 153 krad/h reaction could also result in intramolecular acetal forma-
972 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

tion, producing six-membered rings rather than inter-chain was increased from 0.8 to 5.2 ␮m, which was ascribed to
crosslinking, although no evidence was presented to sub- some pore blocking by polymer chains. In UF of 1 wt% BSA
stantiate this. With fewer OH groups present after reaction of MW 66.5 kDa there was 100% retention of the protein at
with formaldehyde there was a decrease in hydrophilicity 0.34 MPa pressure.
and in the swelling of the membrane. PVA (MW 31–50 kDa, PVA (99% hydrolysed and MW 50 kDa) was crosslinked
98–99% hydrolysed) has been crosslinked with varying with sufficient GA to react with 0.003 mole of the OH
amounts of formaldehyde and the silver nitrate complex groups, and then further reacted under a GA concentration
of the membrane produced tested for the separation of gradient to produce an asymmetry in crosslink density [36].
benzene and cyclohexane [29]. Blend membranes of PVA Excellent selectivities were obtained for creatinine (MW
(MW 125 kDa) and chitosan, with the PVA content rang- 113 Da) over immunoglobulin (150 kDa) relative to results
ing from 20 to 60%, have been made by solution casting with a homogeneous membrane of equivalent crosslinking
and crosslinking with a urea formaldehyde/sulphuric acid to the modified surface network.
mixture [30]. The mixture was made up of 2.5 wt% urea, Lightly crosslinked PVA was reacted with GA in the
2.2 wt% formaldehyde and 2.5 wt% sulphuric acid contain- pores of mixed cellulose ester MF or polysulphone UF
ing 50 wt% aqueous ethanol. Highest permeability was membranes, the amount of reagent again corresponding
obtained with the 60% PVA product, and the highest selec- to only 0.003 mole per mole of PVA repeat unit [37]. The
tivity with the 20% PVA membrane. The membranes were gel-impregnated membranes had potential application in
tested for PV dehydration of isopropanol and tetrahydrofu- biohybrid artificial organs. The most practical configuration
ran. was the hollow fibre mode because of nutrient transport
advantages.
2.7. Glutaraldehyde UF membranes have been made from PVA (>98% hydrol-
ysed and MW 72 kDa) crosslinked with GA in a 1:1 mole
PVA (88% hydrolysed, very high MW) has been used to ratio (enough to react just one of the aldehyde groups with
make a matrix to bind micro (5–10 ␮m) ion-exchange resin the OH groups), and a lipase enzyme covalently bonded
particles into beads of conventional size (300–1200 ␮m), to the surface through the pendant–CHO group [38]. Effi-
with hydrochloric acid as the catalyst and enough glu- cient immobilisation of the enzyme within the pores was
taraldehyde or GA, OHC–(CH2 )3 –CHO, to react with 40% of claimed in testing of the hydrolysis of p-nitrophenyl lau-
the OH groups, or 0.1 mole of GA per mole of PVA repeat unit rate, with the filtration mode avoiding product inhibition
[31]. It should be noted that this is only approximate as GA is of the enzyme.
susceptible to self condensation via an aldol condensation, Macroporous PVA (87.7% hydrolysed and MW 67 kDa)
especially at high pH [32]. crosslinked with GA has been made by carrying out the
The crosslinked matrix was readily permeable to both crosslinking reaction under acidic conditions at a sub-
water and salts. Similarly crosslinked PVA has since been zero temperature of −18 ◦ C overnight [39]. The product,
reported for making membranes for studying diffusive in monolithic form, had interconnected macropores up to
properties and selective separation of flavonoids [33,34]. 150 ␮m in size, with microporosity of the gel walls. The
Membranes for PV of ethanol/water mixtures are material was elastic and sponge-like, with the pore size
reported, made from PVA (98% hydrolysed and MW 75 kDa) decreasing with increase in the concentration of polymer
and GA, with varying degrees of crosslinking [35]. Swelling in the reaction mixture and with decrease in the degree of
in water decreased with increased crosslinking, as did the crosslinking, which varied from 0.05 to 1.0% (w/v) in a reac-
selectivity and permeability of the membrane, but swelling tion mixture that was 5% (w/v) PVA, giving levels of 2–40%
in ethanol was not altered. crosslinking.
Lightly crosslinked PVA (99+% hydrolysed and MW The development of high performance UF membranes
50 kDa) has been deposited on asymmetric regenerated centred on PVA is based on an electrospun nanofibrous
cellulose UF membranes, having a molecular weight cut (150–300 nm diameter) scaffold on which is deposited a
off (MWCO) of 10 kDa, by spin coating a PVA/GA mixture very thin (∼1.8 ␮m) layer of PVA, both crosslinked with GA
onto them, followed by completion of the crosslinking reac- [40,41]. The degree of hydrolysis and MW were optimised
tion [10]. The reaction could well include the OH groups for best performance and mechanical durability of the scaf-
on the cellulose substrate, to give covalent binding of the fold. The range explored was 88–89, 96, 98 and 98–99%
PVA to the support membrane. The extent of crosslinking hydrolysed, with MWs of 13–23, 78 and 85–124 kDa. The
was from 0.005 to 0.01 mole of GA per mole of PVA repeat 96% hydrolysed PVA of highest MW gave the best result.
unit. The lower level gave a layer that was too compress- After crosslinking there was a small volume shrinkage of
ible for thickness measurement. The effect of doubling the <5%, but the porosity was relatively high at >80%. Crosslink-
amount of crosslinking was not significant in fouling and ing of the added layer of PVA with GA, which would have
flux tests with bovine serum albumin (BSA), a result that bonded it to the primary PVA, was varied and the flux and
was attributed to a limiting polymer volume under hydro- rejection characteristics determined. For UF of an oil/water
static pressure. Irreversible fouling was reduced markedly emulsion containing soybean oil (1350 mg/L), a GA/PVA
for membranes with a coating thickness of 3 ␮m or more repeat unit ratio of above 0.06 was needed to achieve a
under hydrostatic pressure. Any protein layer on the surface high rejection (>99.5%), when the permeate flux was then
could be removed readily. Pure water flows were reduced 130 L/m2 h, whereas for the 90% rejection obtained with
as the thickness of the PVA layer was increased, with rel- a ratio of 0.02 the flux was 360 L/m2 h. Thus the size of
ative water fluxes varying from 97 to 47% as the coating the mesh of hydrophilic chains connected by crosslink-
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 973

ing points can be controlled by the degree of crosslinking, time). The crosslinking reaction was hindered by the pres-
with highly crosslinked membrane layers giving much ence of the ␤-cyclodextrin oligomer, which competed for
decreased flux rates. The flux obtained with the optimum the crosslinker. The introduction of charged groups into
membrane was superior to that for a commercial composite a GA-crosslinked PVA membrane increases the selectivity,
UF membrane. Even better results should be possible with but decreases the flux [50].
membranes having a thinner or more hydrophilic coating Varying levels of crosslinking of PVA (94% hydrolysed
layer. and MW 66 kDa) and its effect on PV of acetic acid/water
Similar UF membranes have been made from part mixtures have been tested [51]. Acetone was used as a
acetalised PVA to control their hydrophilicity [42]. The non-solvent for the PVA to prevent it dissolving during the
PVA (99% hydrolysed and MW 1.8 MDa) was reacted with reaction. A minimal amount of 5 vol.% of the GA crosslinker
acetaldehyde under acidic conditions to produce a range gave a water insoluble product, with 10 vol.% giving the
of products having 49–57% acetal groups. The polymers lowest swelling and slowest flux, the separation factor
were cast into membranes from dimethylacetamide solu- decreasing steadily to the highest level used, 30 vol.%. The
tion. The wet membranes were then crosslinked with GA best separation was hence at 5 vol.% when reasonable fluxes
in acid at room temperature. The modified PVA mem- were obtained.
branes were hydrophilic and in the UF of BSA they had PV membranes have been made from PVA (60 wt%)
substantially reduced fouling compared to commercial and poly(allylamine hydrochloride) (35 wt%), crosslinked
membranes. Membrane properties were dependant on the with GA (5 wt%) and shown to give excellent perfor-
degree of acetalisation, the presence of poly(ethylene gly- mance in the dehydration of acetone [52]. The dehydration
col) in the casting solution, and other casting conditions. of ethanol/water mixtures has been explored with PVA
The higher the deacetalisation the more hydrophilic the crosslinked with GA [53]. The cast membranes had two dis-
membrane, giving a better level of water permeation and tinct surfaces, the air-side and glass-side, and performance
good retention of BSA, and with less fouling. was very dependant on which surface faced the feed. Sur-
A composite membrane approach is a three-tiered com- face reconstruction of the membrane did not occur with
posite with a non-woven microporous support, electrospun the glass-side format during PV, so there was no change in
nanofibrous PVA/GA substrate as the middle layer and then water selectivity. The contact angle for the air-side format
a PVA/GA thin coating which contained surface-oxidised decreased with contact time against the feed.
multiwalled carbon nanotubes (CNTs) [43]. The CNTs had Catalytic ceramic disks containing H3 PW12 O40 in the
an outer diameter of 20–40 nm, an inner diameter of pores have been spin-coated with PVA (MW 75 kDa),
5–10 nm and a length of 5–15 ␮m. This unique filtration crosslinked with GA to an undisclosed degree [54]. There
medium gave a high flux rate of up to 330 L/m2 h at a feed was good catalytic stability for the dehydration of butane-
pressure of 0.68 MPa in the UF of an oil/water emulsion, diol to form tetrahydrofuran, with coupled PV to remove
with a rejection of 99.8% and no appreciable fouling after water.
operation for 1 month. An increase in the CNTs in the coat- A summary of relevant published information on PVA
ing layer enhanced the flux rate, which was ascribed to the crosslinked with GA and its applications is given in Table 1.
presence of more effective hydrophilic nanochannels for
water transport through the composite membrane. 2.8. Other aldehydes
PVA/GA membranes containing CNTs are reported,
using ␤-cyclodextrin as a dispersant [44]. They showed Crosslinking PVA with acrolein, CH2 CH–CHO, in the
improved permeate flux and separation efficiency in the form of its bisulphite salt, with the idea of adding the –OH
PV of benzene/cyclohexane mixtures, as well as better groups across the active double bond of acrolein, has been
mechanical properties, than unfilled membranes. The work carried out in the presence of sulphuric acid, followed by
follows similar studies where graphite flakes or carbon curing at 135 ◦ C for 15 min after coating the mixture on a
molecular sieves were introduced into the membrane polysulphone support [24]. In RO experiments, salt rejec-
to give analogous improvements in both flux and selec- tions of 96.5–96.7% from synthetic seawater were achieved
tivity [45–47]. With crystalline flake graphite particles at 10 MPa. A drawback was the compaction of the mem-
of 1–15 ␮m size as the filler, there was a simultaneous brane that was observed at this high operating pressure.
increase in both permeation flux (4-fold) and separation Experiments on PVA membranes made for removal of
factor (6-fold) [48]. The results were attributed to a higher phenol and other components have shown that greater
free volume and more relaxed polymer chain packing rejection is obtained for membranes crosslinked with GA
which could reduce the mass transfer resistance and facil-
itate permeation of benzene. However, excessive filling Table 1
could inhibit permeation. PVA membranes crosslinked with GA.
Membranes for PV of ethanol/water mixtures have
Moles GA per OH Application Reference
been made from similarly crosslinked PVA (99% hydrol-
ysed and MW 75 kDa), but the amount of GA used was 1 Pendant enzyme [38]
0.1 Resin matrix [31]
not specified [49]. Also present in the membrane was ␤- 0.1 PV membrane [51]
cyclodextrin oligomer crosslinked with epichlorohydrin to 0.005–0.1 Chromatography [39]
enhance water permeability and water selectivity (lightly 0.06 UF membrane [41]
crosslinked, 1 h reaction time). More crosslinked versions 0.005–0.01 UF membrane [10]
0.003 Biohybrids [37]
gave increased ethanol permeability as well (8 h reaction
974 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

than those prepared with glyoxal or adipic aldehyde [55]. aldehydes. With trimesic acid (benzene-1,3,5-tricarboxylic
GA has been compared with other aldehydes as crosslink- acid) as the crosslinker, an 88% salt rejection was obtained
ers in the preparation of thin-film composite membranes at a permeate flux of 8 L/m2 h. PVA (>98% hydrolysed and
made by coating a polysulphone UF membrane (MW cut MW 72 kDa) has been crosslinked with malic acid and the
off >15 kDa) with PVA [56]. The PVA (98% hydrolysed and membranes produced tested for the separation of acetic
MWs ranging from 2 to 65 kDa) on the coated membrane acid–water mixtures by PV [62].
was dried and contacted with various crosslinking agents, PVA can be crosslinked effectively with maleic acid,
then heat treated. In RO experiments with 2000 mg/L NaCl cis-HO2 C–CH CH–CO2 H to give membranes useful in the
at 1.7 MPa the salt rejection was as high as 95% for GA, but separation of acetic acid–water mixtures, with 5 mol% of
only 55% for methacrolein crosslinking, with performance the crosslinker resulting in membranes giving higher sepa-
being dependant on the solvent used for the crosslinking ration than membranes based on 10 mol% crosslinker [63].
reaction. Other experiments showed salt rejections of 46% There are more flexible interlinks between PVA chains
for glyoxal, 67% for formaldehyde and 90% for GA. Diffu- than occur with the more crystalline PVA obtained by
sion studies of PVA crosslinked with terephthaldehyde are heat treatment [64]. Crosslinking of PVA (99% hydrol-
reported [57,58]. A similarly crosslinked PVA has been used ysed and MW 75 kDa) with maleic acid or citric acid,
to immobilise enzymes [59]. HO2 C–CH2 –(OH)CH(CO2 H)–CH2 –CO2 H, has been used to
prepare membranes for permeation studies of water-lower
2.9. Di-, tri- and polycarboxylic acids, anhydrides or acid aliphatic alcohol mixtures [65]. Citric acid gave more selec-
chlorides tive behaviour than maleic acid, and the selectivity was
highest for water–isopropanol and water–isobutanol sys-
A study has been made of crosslinking PVA with the tems.
dicarboxylic acids oxalic, malonic and citric acids with PVA (MW 50 kDa) has been crosslinked with
sulphuric acid as catalyst at 90–120 ◦ C [26]. Thin film com- poly(acrylic acid) (MW 2 kDa) by heat treatment at
posite membranes have been made from polysulphone 150 ◦ C for 1 h [66]. The products have been used for PV
membranes, with the PVA as the top layer. Increasing of methanol/methyl tert-butyl ether mixtures, and later
temperatures gave greater salt rejections, which were sig- ethanol–water mixtures [67]. Considerable effort has
nificantly affected by the degree of crosslinking. Increasing gone into producing crosslinked PVA membranes for
the number of carbon atoms in the crosslinker increased both PV separations and fuel cell use, a recent example
the water flux, but the salt rejection was decreased. The being membranes containing sulphonic acid groups [68].
best results were obtained with oxalic acid, with about 95% PVA (MW 89–98 kPa) solutions were mixed with varying
salt rejection at a pressure of 4 MPa, but the water flux amounts (5–30% by weight) of sulphosuccinic acid, cast
was low. Other work on dibasic acids was reviewed, listing and heat treated at 120–130 ◦ C for 2 h. The proton conduc-
membranes with salt rejections of 74–85%. tivity and methanol permeability were very dependant
Asymmetric membranes using PVA (98% hydrolysed on the sulphosuccinic acid content and the crosslinking
and MW 90 kDa) crosslinked with dicarboxylic acids con- temperature.
taining 2–10 carbon atoms have been made for the isolation The crosslinking of PVA (99.5% hydrolysed and MW
of phenol and solvents like methyl isobutyl ketone and 77 kDa) with maleic anhydride in the presence of sulphuric
acetate esters [60]. The reaction was preceded or followed acid catalyst is reported [28]. After drying the membrane
by crosslinking with metal compounds such as chromic was heated at 120 ◦ C for 2 h. Unlike the parallel study
salts. Carbonyl compounds from phenyl ketones to acetone reported above that was carried out with formaldehyde, the
and higher aliphatic ketones were also tested and found to reaction led to mainly inter-chain crosslinking, but there
give membranes of exceptional mechanical properties. In was an analogous decrease in hydrophilicity when greater
tests at 5 MPa, retention of phenol from 2000 mg/L solu- amounts of reagent were used, as OH groups were still con-
tions ranged from 25 to 74%, when malonic and succinic sumed. The water content varied from 2 g/g of membrane
acids were the crosslinkers respectively, while sodium sul- at 1% crosslinking to 1.3 g/g at 15% crosslinking. The prod-
phide rejection was 92–95% (suberic and sebacic acids). uct was tested in PV experiments, where low water fluxes
With ketone crosslinkers the phenol rejection was 56–75% but high separation factors were obtained.
for benzoylacetone and diphenyl ketone. Maleic anhydride has been used to crosslink PVA
Malic acid, HO2 C–CH2 –CH(OH)–CO2 H, has been shown (98.4% hydrolysed, MW 189 kDa), using just 0.05 mole
to be a more effective crosslinking agent than GA or other of the anhydride per mole of PVA [69]. The product
aldehydes and acids in the preparation of thin-film com- was formed on ceramic (␥-Al2 O3 ) hollow fibre supports.
posite membranes made by coating a polysulphone UF It showed a high water flux in the dehydration of 2-
membrane with PVA [56,61]. The PVA used (98% hydrolysed propanol and 1-butanol, plus a simultaneous increase in
and MWs ranging from 2 to 65 kDa) was dried after coat- the separation factor, contradicting the flux-selectivity
ing the support and contacted with the crosslinking agent, paradigm, much as had been observed for nanoparticles of
then heat treated. In RO experiments with 2000 mg/L NaCl fuming silica (<50 nm, 30 wt%) in the stiff-chain polyacety-
at 1.7 MPa the salt rejection was up to 99% and depended lene poly(4-methyl-2-pentyne), which increased both the
on the curing temperature, 120 ◦ C in this case. However, the mixed gas n-butane/methane selectivity and the n-butane
production rate was severely reduced to impractical levels permeability [70]. It was suggested that the presence of
of <5 L/m2 h, versus 29 L/m2 h and 60% rejection when cur- the nanoparticles resulted in a greater polymer/particle
ing was at 70 ◦ C. Rejection was 95% for GA, the best of the interfacial area, disrupting the packing of the rigid and
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 975

bulky polymer chains and enhancing molecular trans- swelling of the PVA [80]. Annealing of the membranes
port. under nitrogen at temperatures of 100, 130 or 160 ◦ C was
A composite membrane has been made from a hol- needed to complete the condensation reaction that intro-
low fibre polysulphone UF membrane by coating it with duced bridging, when higher permselectivity resulted. It
a PVA (95% hydrolysed, MW 1.8 MDa)–sodium alginate was postulated that the crosslinking reaction took place in
(4:1) blend, and using maleic acid in mineral acid as the the non-crystalline parts of the PVA membrane, forming
crosslinking agent [71]. The membrane was tested in the denser non-crystalline regions. Annealing also improved
PV dehydration of aqueous aliphatic alcohol solutions, the selectivity of similar membranes made from poly(vinyl
where high separation factors and permeation fluxes were alcohol-co-acrylic acid) [81]. Similar membranes have been
obtained. Crosslinked PVA/maleic anhydride membranes reported, with the observation that too much crosslinker
have also been fabricated by electrospinning [72]. makes the membrane hydrophobic [82]. Performance was
Poly(vinyl methyl ether) alternating copolymers with enhanced by modifying the membrane by the incorpora-
maleic anhydride have been used to crosslink PVA in inter- tion of chitosan [83]. Poly(ethylene glycol) has also been
polymer membranes [73]. The membrane also contained blended with PVA in such membranes to enhance perme-
sodium poly(styrene sulphonate), but this made the mem- ation properties [84].
brane fragile and weak. UF of erythrosin showed that PVA has been crosslinked with ␥-glycidoxypropyl-
increased crosslinker content gave higher rejections. In RO trimethoxysilane to produce PVA-silica hybrid membranes,
at 40 MPa, salt rejections of up to 98% have been recorded with the aim of improving both permeability and selectiv-
[74]. With ratios of PVA:copolymer of 3.5:1 salt rejections ity in PV [46]. Bridges such as
were >60% with 2000 mg/L NaCl at 2 MPa [75]. Highest
rejection was obtained at pH 11, the lowest at pH 3, as would
be expected from ionisation of the carboxylic acid groups
present in the membranes.
Fumaric acid, trans-HO2 C–CH CH–CO2 H, has been
employed as the crosslinking agent at 0.05 mole per mole
of PVA in multilayer membranes formed on a polyacryloni-
trile support membrane, during which the PVA was filled may be formed as links between the oxygen atoms in the
with various zeolites [76]. The multilayered product was PVA. The permeation flux for benzene from a mixture with
used in the PV of aqueous ethanol. cyclohexane increased from 20.3 g/m2 h for an unfilled PVA
An amic acid has been used as a crosslinker for membrane to 137 g/m2 h for the hybrid membrane, while
PVA [77]. Imidisation at 150◦ gave an improved mem- the separation factor increased from 9.6 to 46.9. This was
brane for PV of aqueous ethanol when there was 12 wt% attributed to an increase in the size and number of both
crosslinker present. More crosslinker showed the reverse network pores and aggregate pores, and an elongation of
effect because of the dispersion of unreacted crosslinker the diffusion path.
within the membrane. A mixture of ␥-glycidoxypropyltrimethoxysilane and
Thin film composite membranes have been prepared by TEOS has been employed in the crosslinking of PVA for
coating a porous polysulphone membrane with an aqueous dehydration of ethylene glycol [85]. The ␥-glycidoxypropyl-
PVA solution containing hydroquinone and a surfactant, trimethoxysilane facilitated the dispersion of silica
and then contacting it with trimesoyl chloride (TMC) in particles throughout the membrane, enhanced the
an organic solvent to give an interfacially polymerised mechanical properties and gave the best selectivity results.
membrane [78]. After drying and heating, a further coat- A more compact crosslinked structure was obtained
ing of PVA solution was applied and the product heated with membranes made from PVA and ␥-mercaptopropyl-
to give a low flux membrane that rejected salt to a high trimethoxysilane [86]. There was no improvement
degree. when the mercaptol groups were oxidised to sul-
TMC has been used to crosslink PVA (99% hydrolysed, phonic acid groups. PVA membranes crosslinked with
MW 133 kDa), in the preparation of PV membranes [79]. ␥-aminopropyltriethoxysilane have been examined in the
The membranes were cast from aqueous PVA solution and separation of ethanol/water mixtures [87]. The incorpo-
the water removed by natural drying. Different volumes of a ration of polysilisesquioxane (PSS) into PVA membranes
2.5% (w/v) TMC/hexane solution were added to the bunded gives hybrid membranes that have increased permse-
membranes which were left to dry again at 25 ◦ C overnight, lectivity and flux at low PSS levels when tested in the
then rinsed with isopropanol and water. The membrane pervaporation dehydration of tetrahydrofuran [88].
thickness was 50–60 ␮m. The crosslinked membranes had Another highly hydroxylated polymer, chitosan or
higher thermal stability than plain PVA film. An appropri- poly(2-amino-2-deoxy-d-glucose), has been used to make
ate degree of crosslinking was found for best performance silica hybrid membranes with ␥-glycidoxypropyltrimet-
in the dehydration of isopropanol. hoxysilane as the crosslinking agent [89,90]. The primary
reaction was through the amino group in chitosan, with
2.10. Alkoxysilanes crosslinking occurring via hydrolysis and condensation
of the methoxy groups, to produce O–Si–O bridges. The
Membranes made from PVA crosslinked with 25 wt% hybrids showed better thermal stability and a low degree
of tetraethoxysilane (TEOS) have been prepared for the of swelling in water. Likewise, in the preparation of PV
PV of aqueous ethanol, with the aim of minimising the membranes for ethanol dehydration, chitosan has been
976 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

crosslinked with 3-aminopropyltriethoxysilane to form tested on 1500 mg/L salt a rejection of 15% was obtained
chitosan-silica hybrid membranes [91]. The hydrophilic- at 0.7 MPa, but a good permeate flux of 14 L/m2 h resulted
ity of the membranes increased with increasing amounts even at the very low test pressure. Divinyl sulphone has
of the crosslinker, reaching a maximum when the con- been used to crosslink PVA coatings of solvent impreg-
tent was 10 wt%. The flux and water selectivity increased nated resins [98]. The PVA (98–99% hydrolysed and MW
as a result. The crosslinking limited the swelling of the 124–186 kDa) was treated with 1 M sodium carbonate prior
polymer. It was claimed that there were strong hydrogen to addition of the crosslinker.
bonds as well as covalent bonds formed, along with a low- PVA has been treated with boric acid and then immersed
ering of crystallinity. Thus the FTIR spectrum showed an in polyvalent ion solutions such as Ba(OH)2 , AlCl3 or
intense absorption band at ∼1072 cm−1 , indicating that a KCr(SO4 )2 [6]. The products are chemically stable and show
reaction took place between the OH groups in the polymer product fluxes of 6.3–10.4 L/m2 h in RO at 5 MPa. The reac-
and the silanol groups in the crosslinker, to form C–O–Si tion of boric acid or its salts with PVA has been thoroughly
bonds. If the second ethoxy group reacts with the OH in explored [99], and although they do not crosslink PVA,
another chain, bridges can occur. The crosslinking makes gelation does occur in the presence of cations. Complex
the polymer chains in the amorphous regions more com- formation with metal ions has also been utilised [100],
pact, resulting in less space for species to permeate through followed by crosslinking with acidic glutaraldehyde. The
the membrane and making the resistance higher for the diffusion properties of these membranes have been tested
larger species. with theophylline and vitamin B12 .
Gaseous 1,2-dibromoethane was the crosslinker chosen
2.11. Other chemical crosslinking agents for the polymer matrix of interpenetrating polymer net-
works forming an ion-exchange membrane from PVA (99%
Composite RO membranes have been constructed from hydrolysed and MW 124–186 kDa) at 130 ◦ C [101].
Millipore filters with an ultrathin layer of PVA crosslinked Epichlorohydrin has been employed in the crosslinking
with toluene di-isocyanate (TDI) [92]. Operation at 10 MPa of hydrolysed starch-g-polyacrylonitrile/PVA blend films
yielded salt rejections of 98% and a water flux of 21 L/m2 h. by reaction at pH 10 at 40 ◦ C for 2 h [102]. The hydrol-
TDI has also been used to surface modify a PVA UF mem- ysed graft polymer contains carboxylic acid and amide
brane that had been crosslinked with GA, using 0.005 moles groups which can be linked to the OH groups of PVA by
of GA per OH group [93]. Compared to the unmodified the crosslinker. Tests on RO behaviour have been made
membrane the new product had a 4.5-fold improvement [103].
in selectivity for lysozyme (MW 14.1 kDa) over BSA (MW It is interesting to note that macroporous PVA mem-
66.5 kDa). The barrier layer provided a higher level of size branes can be made by an electrospinning technique
exclusion. There was a small reduction in lysozyme flux to from a solution of PVA (98% hydrolysed and MW 77 kDa)
76% of the original value. Mixed cellulose ester MF mem- containing an ethoxylated nonylphenol surfactant [104].
branes of 0.3 ␮m pore size have been impregnated with Stabilising the film against dissolution in water can be
such gels to provide a support and give a thinner, less frag- achieved by crosslinking with bifunctional reagents such as
ile membrane [94]. A di-isocyanate terminated oligoether sodium borate, boric acid or GA, as well as via irradiation
made from tri(ethylene glycol) and TDI has been used to or alternating freeze–thaw conditions.
crosslink cellulose that had grafts of poly(ethylene oxide) PVA has been crosslinked with 1,1,1-tris(hydroxy-
[95]. The products have application as separators in lithium methyl)propane by heating at 150 ◦ C for 30 min in the pres-
ion batteries. ence of an acid catalyst [105]. An insoluble film is formed,
In what amounts to crosslinking with glycidyl acry- the porosity of which can be enhanced by incorporating
late, macromers have been made by reacting PVA sodium poly(styrene sulphonate) within the film as an
(99% hydrolysed, MW 13–23 kDa) with the epoxy com- interpenetrating polymer. Swelling in water ranged from
pound under acidic conditions to form the PVA ether 50 to 60%, and erythrosin rejection in UF operation from 81
containing pendant acrylate groups [96]. A photoinitia- to 88%.
tor, 2-hydroxy-1-[4-(hydroxyethoxy)phenol]-2-methyl-1- A recent article is devoted to the preparation of adsor-
propanone, was added to aqueous solutions of the polymer bent membranes from chitosan. The membranes were
which were polymerised by exposure to ultraviolet light. formed with a macro open pore structure using salt par-
The degree of swelling in water was a function of the reac- ticles of size 150–250 ␮m, and genipin as the crosslinking
tant concentration, ranging from 5.5 to 15 to 18 as the agent [106]. Genipin is an aglycone derived from an iridoid
concentration was changed from 50 to 30 to 20 wt%, with glycoside present in fruit of Gardenia jasmindides Ellis. It
complete reaction occurring at all levels. The more water is much less toxic than other chemical crosslinkers, and is
present in the reaction mixture, the greater the amount of used for pharmaceutical purposes. The reaction is carried
intramolecular reaction of the pendant double bonds. It was out in 5 vol.% acetic acid solution at ambient tempera-
found that the resulting network in the best case was looser ture for 10 h, and then evaporated at 50◦ . The membrane
than that obtained for networks crosslinked with a compa- swelling in water ranged from 2.5 to 4.2.
rable amount of GA at the 2 and 7 mole% levels. Physical
crosslinking via the introduction of crystallinity gave even 3. Grafting of PVA onto support membranes
looser networks.
The crosslinking of PVA with divinyl sulphone under PVA, and also hydroxyethyl cellulose and dextran, have
alkaline conditions at 70 ◦ C has been reported [97]. When been covalently linked to polyamide membranes that had
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 977

been activated with bis-oxirane (1,6-hexanediol diglycidyl of 26–23.5 L/m2 h, which was 91% of that from the totally
ether), then a dye immobilised on the surface to form dye- hydrophobic membrane. The wetting problem facing tra-
affinity membranes [107]. The membranes were used to ditional MD with the hydrophobic membrane was avoided
separate an enzyme from fermentation broth by retention. by incorporating the hydrophilic layer. There can be serious
The permeability of the membranes was reduced by about problems with membrane wetting of hydrophobic mem-
half compared to that of the original polyamide membrane branes, causing MD to slow down and, in the worst cases,
because of the extension of polymer coils into the pores. The being brought to a halt.
presence of cell debris caused only a slight decrease in per-
formance, indicating minimal attachment of these particles
to the hydrophilic surface. 4. Commercial PVA membranes
A simple method of grafting PVA onto porous hydropho-
bic membranes such as polyethylene, polypropylene or There is limited information on commercial membranes
poly(vinylidene fluoride) involves pretreating the mem- in the open literature. NF composite membranes have
brane with chromic acid solution then dipping it in a PVA been prepared by coating a mixture of PVA and a reactive
(99+% hydrolysed and MW 85–146 kDa) solution that had dye onto a porous polypropylene membrane, followed by
been crosslinked with maleic acid or GA [108]. In the case crosslinking by heating [112,113]. Stable to pH 12 at 60 ◦ C,
of GA the amount of crosslinker was 0.0025– 0.01 moles the membranes give 98% rejection of ionic dyes with MWCO
per PVA repeating unit. Some bonding of the PVA via the of ≥700 Da at 2–3 MPa. They have been used industrially in
crosslinker to –OH groups introduced to the membrane sur- the desalting of dye liquors.
face must take place. The membranes were used for osmotic To overcome slow fluxes, porous supporting membranes
distillation of oily feeds. have been coated with thin PVA films crosslinked with a
Grafting of proteins onto crosslinked PVA surfaces by dialdehyde, using phosphoric acid as catalyst and pore for-
the use of oxidised starch has been explored [109]. The mer at 110 ◦ C [113,114]. The resulting NF membranes have
coupling is based on Schiff base formation between the found application in copper recovery from alkaline plating
amino groups of proteins and the aldehyde groups of oxi- baths. The membrane was stable to pH 13. Rejection of NaCl
dised starches that have been grafted onto the substrate was 20% and MgSO4 , 85%.
membrane through acetalisation of the aldehyde of starch Low pressure RO membranes with high salt rejection
with the –OH groups of the substrate polymer. Another have been made from polysulphone UF substrates by form-
way of doing this would be to react the substrate polymer, ing a thin skin of PVA on the surface [115]. No preparative
PVA as the prime example, with sufficient dialdehyde to details are given, but the PVA must be crosslinked to ensure
leave a pendant –CHO group at each site. The reverse situa- insolubility. Two such membranes made by the Nitto Elec-
tion would be the reaction of the dialdehyde with terminal tric Industrial Company (now Nitto Denko, Kusatsu, Japan),
amino groups in a polyamide membrane to produce pen- used in spiral wound mode, have been designated NTR-
dant –CHO groups, which in turn would be reacted with 739HF and NTR-729HF. They can desalt brackish water at
PVA in a grafting process. 1 MPa pressure to give salt rejections of 95.5 and 91.5%
Commercial RO and NF membranes that have been respectively, at fluxes of 33 and 54 L/m2 h. An important
coated with PVA are less fouled in the treatment of dye feature is their resistance to chlorine attack.
house wastewater than uncoated membranes [110]. This These commercial PVA membranes have been used to
was ascribed to the reduction of both the surface nega- separate the constituents of a lipase hydrolysate of high-
tive charge and the surface roughness that accompanied oleic sunflower oil after dilution with organic solvents
the coating: a neutral and smoother surface was signifi- [116]. The pressure used was 4 MPa at 40 ◦ C, when the mem-
cantly beneficial. Passage of NaCl and CaCl2 was slightly branes performed as well as a polyamide RO membrane in
reduced for the NF membrane, and slightly increased for terms of rejection or selectivity, but not as well as cellulose
the RO membrane. acetate membranes. The RO performance of the two PVA
A membrane has been made from PVA blended with systems was quoted as 93 and 60% salt rejection respec-
poly(ethylene glycol) or PEG, and crosslinked with a tively by the supplier. Under the same conditions the salt
dialdehyde [111]. It was investigated as a modified com- rejections of a cellulose acetate and a polyamide membrane
posite membrane for membrane distillation (MD), where were 98 and 99.5% respectively.
normally water vapor is transported through a porous A similar product has been reported that has an aro-
hydrophobic membrane. A PVDF membrane of 0.2 ␮m pore matic polyamide RO membrane as the substrate [117]. After
size was used as a substrate, pretreating one side of it coating with the PVA, the composite membrane was heat
with a potassium dichromate/sulphuric acid solution and treated at 130 ◦ C for 5 min, resulting in a 0.1 ␮m thick
casting a PVA (MW 2 kDa)/PEG (MW 10 kDa) solution that layer being formed on the original membrane. For both the
was then crosslinked with the dialdehyde, which would coated and uncoated membranes the salt rejection from a
bond the PVA to the treated membrane. The hydrophilic 1500 mg/L NaCl solution was 99.5–99.7% at an operating
layer obtained was placed in contact with the salt solu- pressure of 150 kPa and a flux of 4–25 L/m2 h, showing that
tion. In desalting 3.5% NaCl, 91% of the flux obtained with the main salt barrier was provided by the supporting aro-
a purely hydrophobic membrane was achieved with a sep- matic polyamide membrane. High anti-fouling properties
aration efficiency above 99% when the temperature of the were claimed for the coated product.
feed was 70 ◦ C. The lower the cooling temperature, which Low fluxes are claimed to be the reason why PVA mem-
varied from 12 to 25 ◦ C, the higher the flux, with a range branes are not in large-scale commercial use [118].
978 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

Table 2 Table 4
Attributes of PVA membranes. RO results for PVA membranes crosslinked by different means.

Positive Negative Crosslinker Pressure NaCl Rejn. Flux Reference


(MPa) (%) (L/m2 h)
Excellent hydrophilicity High degree of swelling
Permeability to water Permeability to ions Heat treatment 5.2 26 – [15]
Good mechanical properties Compaction under pressure Heat treatmenta 8 93 28 [16]
Thermal resistance Low flux when highly crosslinked ␥-Irradiation 10 20–50 – [20]
Resistance to chemicals Acid-catalysed dehydration 2 70–85 29–33 [4]
Anti-fouling potential Persulphate 2 45 42 [22]
Low operating pressure Formaldehyde 4 93–97 – [23]
Film forming ability Formaldehyde 10 90 – [24]
Formaldehyde 5 25–28 35-54 [25]
Formaldehyde 1.7 67 – [56]
5. Conclusions Glutaraldehyde 4 85 – [26]
Glutaraldehyde 1.7 99 <5 [56]
Glyoxal 1.7 46 – [56]
The main characteristics of PVA membranes are sum- Oxalic acid 4 95 – [26]
marised in Table 2, and the types of crosslinking agents Malic acid 1.7 60–99 – [65]
that have been used on PVA are listed in Table 3. A fur- Trimesic acid 1.7 88 8 [56]
Maleic anhydride 40 98 – [75]
ther point worth considering is the toxicity of the reagents
copolymer
or the by-products formed during the crosslinking. This Maleic anhydride 2 >60% – [76]
applies particularly to formaldehyde, which in the presence copolymer
of hydrochloric acid can form chloromethyl ether, the use Toluene diisocyanate 10 98 21 [92]
of which is highly restricted as it poses a significant hazard. Acrolein 10 97 – [24]
Methacrolein 1.7 55 – [56]
Other crosslinkers such as toluene diisocyanate are unlikely
Divinyl sulphone 0.7 15 14 [98]
to be approved for potable water production. Nitti Denko NTR-739HFb 1.0 95.5 33 [115]
The beneficial properties relevant to membrane opera- Nitto Denko productc 0.15 99.7 4–25 [117]
tion are many, outweighing the detrimental ones. Arising a
A 3:2 PVA/poly(styrene sulphonic acid) composite.
from the superior hydrophilicity, the disadvantages can be b
PVA on polysulphone support, crosslinker unknown.
overcome by crosslinking reactions through some of the OH c
PVA on aromatic polyamide support.
groups responsible for the hydrophilicity. However, this is
really counterproductive: what is really needed is a net-
water permeability and permeate flux rate. This arises
work that provides a tight restraining without serious loss
because of the tighter network, but there is a decrease in
of hydrophilic behaviour.
hydrophilicity as some of the OH groups are consumed by
PVA membranes crosslinked by a variety of means have
reaction with the crosslinking agent. There is hence an opti-
been tested for RO salt rejection as detailed in Table 4.
mum degree of crosslinking for each bridging system. To
GA is a more effective crosslinking agent than formalde-
overcome the possible low fluxes associated with a high
hyde or glycidyl acrylate, which in turn gives a less swollen
degree of crosslinking, thin film composite structures are
product than that obtained by increasing the crystallinity
appropriate, as have been developed in early commercial
by heating. TDI and acrolein give similar results, but at an
PVA membranes.
extremely high applied pressure. Crosslinking with maleic
The high swelling behaviour can be countered also by
anhydride/vinyl methyl ether copolymers gives as good a
forming the active PVA component inside the pores of an
result, but at even higher pressure.
MF membrane. Crosslinked PVA inside such membranes
The degree of crosslinking is not always specified. Good
would have its swelling suppressed, and could function as
crosslinking is crucial for performance as highly crosslinked
the active RO material; by merely coating the pore walls,
membranes will have a higher salt rejection, but a lower
leaving porosity, MF or UF membranes could be prepared. In
both situations some grafting to the host membrane would
Table 3 be beneficial.
Crosslinkers used on PVA.

Freeze–thaw treatment Malic acid References


Heat treatment Malonic acid
Acid-catalysed dehydration Fumaric acid
[1] Carroll T, Booker NA, Meier-Haack J. Polyelectrolyte-grafted micro-
␥-Irradiation Poly(acrylic acid) filtration membranes to control fouling by natural organic matter
Persulphate treatment Trimesic acid in drinking water. J Membr Sci 2002;203:3–13.
Formaldehyde Trimesoyl chloride [2] Knoel T, Safarik J, Cormack T, Riley R, Lin SW, Ridgway H.
Glutaraldehyde Toluene diisocyanate Biofouling potentials of microporous polysulfone membranes coan-
Glyoxal Glycidyl acrylate taining a sulfonated polyether-ethersulfone/polethersulfone block
Terephthaldehyde Divinyl sulphone copoloymer: correlation of membrane surface properties with bac-
Acrolein & methacrolein Boric acid terial attachment. J Membr Sci 1999;157:117–38.
Urea formaldehyde/H2 SO4 1,2-Dibromoethane [3] Finch CA. Chemical reactions and stereochemistry of polyvinyl alco-
Citric acid Tetraethoxysilane hol. In: Finch CA, editor. Polyvinyl alcohol–developments. New York:
Maleic acid & anhydride ␥-Glycidoxypropyltrimethoxysilane Wiley; 1992. p. 269–312.
Maleic anhydride ␥-Mercaptopropyltrimethoxysilane [4] Immelman E, Sanderson RD, Jacobs EP, van Reenan AJ. Poly(vinyl
alcohol) gel sub-layers for reverse osmosis membranes. I. Insol-
copolymers with vinyl
ubilisation by acid-catalysed dehydration. J Appl Polym Sci
methyl ether
1993;50:1013–34.
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 979

[5] Mallapragada SK, Chin SF. Multilayered semicrystalline polymeric pervaporation dehydration of isopropanol and tetrahydrofuran. J
controlled release systems. In: McCulloch I, Shalaby SW, editors. Appl Polym Sci 2006;103:1918–26.
Tailored polymeric materials for controlled delivery systems. ACS [31] Bolto BA, Eppinger KH, Macpherson AS, Siudak R, Weiss DE,
Symposium Series 709. Washington: American Chemical Society; Willis D. An ion-exchange process with thermal regeneration IX.
1998. p. 176–84. A new type of rapidly reacting ion-exchange resin. Desalination
[6] Peter S, Hese N, Stefan R. Phenol-selective, highly resistant RO mem- 1973;13:269–85.
branes made from PVA for the purification of toxic industrial wastes. [32] Tashima T, Imai M, Kuroda Y, Yagi S, Nakagawa T. Structure of a
Desalination 1976;9:161–7. new oligomer of glutaraldehyde produced by aldol condensation
[7] Katz MG, Wydeven T. Poly(vinyl alcohol membranes for reverse reaction. J Organic Chem 1991;56:694–7.
osmosis. In: Turbak AF, editor. Synthetic membranes. ACS Sympo- [33] Reinhart CT, Peppas NA. Solute diffusion in swollen membranes.
sium Series 153. Washington: American Chemical Society; 1981. p. Part II. Influence of crosslinking on diffusive properties. J Membr
383–98. Sci 1984;18:227–39.
[8] Kim KJ, Fane AG, Fell CJD. The performance of ultrafiltration mem- [34] Yoshizuka K, Ohta H, Inuoe K, Kitazaki H, Ishimaru M. Selective sep-
branes pretreated by polymers. Desalination 1988;70:229–49. aration of flavonoids with a polyvinyl alcohol membrane. J Membr
[9] Brink LES, Romijn DJ. Reducing the protein fouling of polysulfone Sci 1996;118:41–8.
surfaces and polysulfone ultrafiltration membranes: optimisation [35] Kim K-J, Lee S-B, Han N-W. Effects of degree of crosslink-
of the type of presorbed layer. Desalination 1990;78:209–33. ing on properties of poly(vinyl alcohol) membranes. Polym J
[10] Li RH, Barbari TA. Performance of poly(vinyl alcohol) thin-gel com- 1993;25:1293–302.
posite ultrafiltration membranes. J Membr Sci 1995;105:71–8. [36] Dai WS, Barbari TA. Hydrogel membranes with mesh size asym-
[11] Amanda A, Mallapragada SK. Comparison of protein fouling on metry based on the gradient crosslinking of poly(vinyl alcohol). J
heat-treated poly(vinyl alcohol), poly(ether sulfone) and regener- Membr Sci 1999;156:67–79.
ated cellulose membranes using diffuse reflectance infrared Fourier [37] Dai WS, Barbari TA. Gel-impregnated pore membranes with
transform spectroscopy. Biotechnol Prog 2001;17:917–23. mesh-size asymmetry for biohybrid artificial organs. Biomaterials
[12] Hickey AS, Peppas NA. Mesh size and diffusive characteristics of 2000;21:1363–71.
semicrystalline poly(vinyl alcohol) membranes prepared by freez- [38] Djennad M, Benachour D, Berger H, Schomäcker R. Crosslinked
ing/thawing techniques. J Membr Sci 1995;107:229–37. poly(vinyl alcohol) ultrafiltration membranes: Synthesis, char-
[13] Ofstead RF, Poser CI. Semicrystalline poly (vinyl alcohol) hydro- acterisation, the use of enzyme immobilisation. Eng Life Sci
gels. In: Glass AG, editor. Polymers in aqueous media: Performance 2003;3:446–52.
through association. Advances in Chemistry Series 223. Washing- [39] Plieva FM, Karlsson M, Aguilar M-R, Gomez D, Mikhalovsky IY,
ton: American Chemical Society; 1989. p. 61–72. Mattiasson B. Pore structure of macroporous monolithic cryogels
[14] Katz MG, Wydeven T. Selective permeability of PVA membranes. II. prepared from poly(vinyl alcohol). J Appl Polym Sci 2006;100:
Heat-treated membranes. J Appl Polym Sci 1982;27:79–87. 1057–66.
[15] Reid CE, Breton EJ. Water and ion flow across cellulosic membranes. [40] Koski A, Yim K, Shivkumar S. Effect of molecular weight on fibrous
J Appl Polym Sci 1959;1:133–43. PVA produced by electrospinning. Mater Lett 2004;58:493–7.
[16] Koyama K, Okada M, Nishimura M. An interpolymer anionic [41] Wang X, Fang D, Yoon K, Hsiao BS, Chu B. High performance
composite reverse osmosis membrane derived from poly(vinyl alco- ultrafiltration composite membranes based on poly(vinyl alcohol)
hol) and poly(styrene sulphonic acid). J Appl Polym Sci 1982;27: hydrogel coating on crosslinked nanofibrous poly(vinyl alcohol)
2783–9. scaffold. J Membr Sci 2006;278:261–8.
[17] Yu X, Wang A, Cao S. Water-vapor permeability of polyvinyl alcohol [42] Zhang Y, Li H, Li H, Li R, Xiao C. Preparation and characterisation of
films. Desalination 1987;62:293–7. modified poly(vinyl alcohol) ultrafiltration membranes. Desalina-
[18] Wang Y-L, Yang H, Xu Z-L. Influence of post-treatments on the prop- tion 2006;192:214–23.
erties of porous poly(vinyl alcohol) membranes. J Appl Polym Sci [43] Wang X, Chen X, Yoon K, Fang D, Hsiao BS, Chu B. High flux fil-
2008;107:1423–9. tration medium based on nanofibrous substrate with hydrophilic
[19] Nitto Electric Industrial Co. Ltd. Composite semipermeable mem- nanocomposite coating. Environ Sci Technol 2005;39:7684–91.
brane production by forming a membrane of PVA or partially [44] Peng F, Hu C, Jiang Z. Novel poly(vinyl alcohol)/carbon nan-
saponified PVAC on a porous support and treating with a otube hybrid membranes for pervaporation separation of ben-
water-soluble polymer containing carboxyl groups. Nitto Electric zene/cyclohexane mixtures. J Membr Sci 2007;297:236–42.
Industrial Co. Ltd. Jap. Patent JP59,69,107 (1984). [45] Peng F, Lu L, Sun H, Hu C, Wu H, Jiang Z. Significant increase
[20] Katz MG, Wydeven T. Selective permeability of PVA mem- of permeation flux and selectivity of poly(vinyl alcohol) mem-
branes. I. Radiation-crosslinked membranes. J Appl Polym Sci branes by incorporation of crystalline flake graphite. J Membr Sci
1981;26:2935–46. 2005;259:65–73.
[21] Sanderson RD, Immelman E, Bezuidenhout D, Jacobs EP, van Reenen [46] Peng F, Lu L, Sun H, Wang Y, Liu J, Jiang Z. Hybrid organic–inorganic
AJ. Polyvinyl alcohol and modified polyvinyl alcohol reverse osmosis membrane: solving the tradeoff between permeability and selectiv-
membranes. Desalination 1993;90:15–29. ity. Chem Mater 2005;17:6790–6.
[22] Immelman E, Bezuidenhout D, Sanderson RD, Jacobs EP, van [47] Sun H, Lu L, Peng F, Wu H, Jiang Z. Pervaporation of ben-
Reenan AJ. Poly(vinyl alcohol) gel sub-layers for reverse osmosis zene/cyclohexane mixtures through CMS-filled poly(vinyl alcohol)
membranes. III. Insolubilisation by crosslinking with potassium per- membranes. Sep Purif Technol 2006;52:203–8.
oxydisulphate. Desalination 1993;94:115–32. [48] Peng F, Hu C, Jiang Z. Organic–inorganic hybrid membranes with
[23] Chen CT, Chang YJ, Chen MC, Tobolsky AV. Formalised poly(vinyl simultaneously enhanced flux and selectivity. Ind Eng Chem Res
alcohol) membranes for reverse osmosis. J Appl Polym Sci 2007;46:2544–9.
1973;17:789–96. [49] Yamasaki A, Mizoguchi K. Preparation of PVA membranes con-
[24] Cadotte JE, Cobian KE, Forester RH, Petersen RJ. Continued evalua- taining ␤-cyclodextrin oligomer (PVA-CD membrane) and their
tion of in-situ formed condensation polymers for reverse osmosis pervaporation characteristics for ethanol/water mixtures. J Appl
membranes. NTIS Report No. PB-253193, 1976. Polym Sci 1994;51:2057–62.
[25] Chang HN. Reverse osmosis separation of inorganic salts using [50] Praptowidodo VS. Influence of swelling on water transport through
polyvinyl alcohol membranes. Desalination 1982;42:63–77. PVA-based membrane. J Mol Struct 2005;739:207–12.
[26] Jian S, Ming SX. Crosslinked PVA-PS thin-film composite membrane [51] Yeom C-K, Lee K-H. Pervaporation separation of water-acetic acid
for reverse osmosis. Desalination 1987;62:395–403. mixtures through poly(vinyl alcohol) membranes crosslinked with
[27] Durmaz-Hilmioglu N, Yildirim AE, Sakaoglu AS, Tulbentei S. Acetic glutaraldehyde. J Membr Sci 1996;109:257–65.
acid dehydration by pervaporation. Chem Eng Proc 2001;40:263–7. [52] Namboodiri VV, Ponangi R, Vane LM. A novel hydrophilic polymer
[28] Han B, Li J, Chen C, Xu C, Wickramasinghe SR. Effects of degree of membrane for the dehydration of organic solvents. Eur Polym J
formaldehyde acetal treatment and maleic acid crosslinking on sol- 2006;42:3390–3.
ubility and diffusivity of water in PVA membranes. Trans I Chem Eng [53] Li G, Zhang W, Yang J, Wang X. Time-dependence of pervaporation
2003;81A:1385–92. performance for the separation of ethanol/water mixtures through
[29] Bryant DL, Noble RD, Koval CA. Facilitated transport of benzene and poly(vinyl alcohol) membrane. J Coll Interface Sci 2007;306:337–44.
cyclohexane with poly(vinyl alcohol)-AgNO3 membranes. J Membr [54] Liu Q-L, Li Q-B. Membrane of PVA coated on porous catalytic ceramic
Sci 1997;127:161–70. disks supported H3 PW12 O40 . J Membr Sci 2002;202:89–95.
[30] Krishna Rao KSV, Subha MCS, Sairam M, Mallikarjuna NN, Aminab- [55] Peter S, Stefan R. Chemically resistant asymmetric membranes made
havi TM. Blend membranes of chitosan and poly(vinyl alcohol) in from PVA for the separation of organic solvents and phenols from
980 B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981

aqueous solution. In: Turbak AF, editor. Synthetic membranes. ACS [78] Toray Industries Inc. Semipermeable composite membrane–
Symposium Series 153. Washington: American Chemical Society; comprising porous support of crosslinked polyester and poly-
1981. p. 281–91. condensed prod. of aromatic poly:phenol and aromatic poly:acid
[56] Lang K, Sourirajan S, Matsuura T, Chowdhury G. A study on the chloride Toray Industries Inc. Jap. Patent JP58,89,910 (1983).
preparation of polyvinyl alcohol thin-film composite membranes [79] Xiao S, Huang RYM, Feng X. Preparation and properties of trimesoyl
and reverse osmosis testing. Desalination 1996;104:185–96. chloride crosslinked poly(vinyl alcohol) membranes for pervapora-
[57] Ebra-Lima OM, Paul DR. Hydraulic permeation of liquids through tion dehydration of isopropanol. J Membr Sci 2006;286:245–54.
swollen polymeric networks. I. Poly(vinyl alcohol)–water. J Appl [80] Uragami T, Okazaki K, Matsugi H, Miyata T. Structure and
Polym Sci 1975;19:1381–6. permeation characteristics of an aqueous ethanol solution of
[58] Braun D, Walter E. Intra- und intermolekulare Vernetzung von organic–inorganic hybrid membranes composed of poly(vinyl alco-
Polyvinylalkohol. Colloid Polym Sci 1976;254:396–9. hol) and tetraethoxysilane. Macromolecules 2002;35:9156–63.
[59] Manecke G, Polakowski D. Some carriers for the immobilisation of [81] Uragami T, Matsugi H, Miyata T. Pervaporation characteristics
enzymes based on derivatised poly(vinyl alcohol) and on copoly- of organic-inorganic hybrid membranes composed of poly(vinyl
mers of methacrylates with different spacer lengths. J Chromatogr alcohol-co-acrylic acid) and tetraethoxysilane for water ethanol
1981;215:13–24. separation. Macromolecules 2005;38:8440–6.
[60] Peter S, Stefan R. The separation of organic solvents and phenols [82] Kulkarni SS, Kittur AA, Aralaguppi MI, Kariduraganavar MY. Syn-
from aqueous solutions by means of chemically stable asymmetric thesis and characterisation of hybrid membranes using poly(vinyl
membranes derived from PVA. In: Proc 7th Int Symp Fresh Water alcohol) and tetraorthosilicate for the pervaporation separation of
from the Sea. 1980. p. 197–209. water-isopropanol mixtures. J Appl Polym Sci 2004;94:1304–15.
[61] Lang K, Matsuura T, Chowdhury G, Sourirajan S. Preparation and [83] Kulkarni SS, Tambe SM, Kittur AA, Kariduraganavar MY. Modifica-
testing of polyvinyl alcohol composite membranes for reverse tion of tetraorthosilicate crosslinked poly(vinyl alcohol) membrane
osmosis. Can J Chem Eng 1995;73:686–92. using chitosan and its application to the pervaporation separa-
[62] Işiklan N, Şanh O. Separation characteristics of acetic acid-water tion of water-isopropanol mixtures. J Appl Polym Sci 2006;99:
mixtures by pervaporation using poly(vinyl alcohol) mem- 1380–9.
branes modified with malic acid. Chem Eng Proc 2005;44: [84] Ye LY, Liu QL, Zhang QG, Zhu AM, Zhou GB. Pervaporation
1019–27. characteristics and structure of poly(vinyl alcohol)/poly(ethylene
[63] Huang RYM, Rhim JW. Modification of poly(vinyl alcohol) using glycol)/tetraethoxysilane hybrid membranes. J Appl Polym Sci
maleic acid and its application to the separation of acetic 2007;105:3640–8.
acid–water mixtures by the pervaporation technique. Polym Int [85] Guo R, Hu C, Pan F, Wu H, Jiang Z. PVA-GPTMS/TEOS hybrid per-
1993;30:129–35. vaporation membrane for dehydration of ethylene glycol aqueous
[64] Gohil JM, Bhattacharya A, Ray P. Studies on the crosslinking of solution. J Membr Sci 2006;281:454–62.
poly(vinyl alcohol). J Polym Res 2006;13:181–9. [86] Guo R, Ma X, Hu C, Jiang Z. Novel PVA-silica nanocomposite mem-
[65] Burshe MC, Sawant SB, Joshi JB, Pangarkar VG. Sorption and per- brane for pervaporative dehydration of ethylene glycol aqueous
meation of binary water-alcohol systems through PVA membranes solution. Polymer 2007;48:2939–45.
crosslinked with multifunctional crosslinking agents. Sep Purif [87] Zhang QG, Liu QL, Jiang ZY, Chen Y. Anti-trade-of in dehydration
Technol 1997;12:145–56. of ethanol by novel PVA/APTEOS hybrid membranes. J Membr Sci
[66] Rhim J-W, Kim Y-K. Pervaporation separation of MTBE-methanol 2007;287:237–45.
mixtures using crosslinked PVA membranes. J Appl Polym Sci [88] Zhang QG, Liu QL, Shi FF, Xiong Y. Structure and permeation of
2000;75:1699–707. organic-inorganic hybrid membranes composed poly(vinyl alcohol)
[67] Lee K-H, Kim H-K, Rhim J-W. Pervaporation separation of and polysilisesquioxane. J Mater Chem 2008;18:4646–53.
binary organic-aqueous liquid mixtures using crosslinked PVA [89] Liu Y-L, Su Y-H, Lai J-Y. In situ crosslinking of chitosan
membranes. III. Ethanol–water mixtures. J Appl Polym Sci and formation of chitosan-silica hybrid membranes using ␥-
2003;58:1707–12. glycidoxypropyltrimethoxysilane as a crosslinking agent. Polymer
[68] Rhim J-W, Park HB, Lee C-S, Jun J-H, Kim DS, Lee YM. Crosslinked 2004;45:6831–7.
poly(vinyl alcohol) membranes containing sulfonic acid group: pro- [90] Liu Y-L, Su Y-H, Lee K-R, Lai J-Y. Crosslinked organic-inorganic
ton and methanol transport through membranes. J Membr Sci hybrid chitosan membranes for pervaporation dehydration of
2004;238:143–51. isopropanol-water mixtures with a long term stability. J Membr Sci
[69] Peters TA, Poeth CHS, Benes NE, Buijs HCWM, Vercauteren FF, 2005;251:233–8.
Keurentjes JTF. Ceramic-supported thin PVA pervaporation mem- [91] Chen JH, Liu QL, Zhang XH, Zhang QG. Pervaporation and
branes combining high flux and high selectivity; contradicting the characterisation of chitosan membranes cross-linked by 3-
flux-selectivity paradigm. J Membr Sci 2006;276:42–50. aminopropyltriethoxysilane. J Membr Sci 2007;292:125–32.
[70] Merkel TC, Freeman BD, Spontak RJ, He Z, Pinnau I, Meakin P, et [92] Dick R, Nicolas L. Membranes composites preparees a partir d’alcool
al. Sorption, transport and structural evidence for enhanced free polyvinylique et de diisocyanate de toluene destinees a l’osmose
volume in poly(4-methyl-2-pentyne)/fumed silica nanocomposite inverse. Desalination 1975;17:239–55.
membranes. Chem Mater 2003;15:109–23. [93] Li RH, Barbari TA. Protein transport through membranes based on
[71] Dong YQ, Zhang L, Shen JN, Song MY, Chen HL. Preparation of toluene diisocyanate surface-modified poly(vinyl alcohol) gels. J
poly(vinyl alcohol)-sodium alginate hollow-fibre composite mem- Membr Sci 1994;88:115–25.
branes and pervaporation dehydration characterisation of aqueous [94] Li RH, Barbari TA. Characterisation and mechanical support of
alcohol mixtures. Desalination 2006;193:202–10. asymmetric hydrogel membranes based on the interfacial crosslink-
[72] Yang E, Qin X, Wang S. Electrospun crosslinked polyvinyl alcohol ing of poly(vinyl alcohol) with toluene diisocyanate. J Membr Sci
membrane. Mater Lett 2008;62:3555–7. 1996;111:115–22.
[73] Gryte CC, Chen J, Kevorkian V, Gregor HP. Hydrophilic interpoly- [95] Chelmecki M, Meyer WH, Wegner G. Effect of crosslinking on poly-
mer membranes from crosslinked blends of poly(vinyl alcohol), mer electrolytes based on cellulose. J Appl Polym Sci 2007;105:25–9.
poly(styrene sodium sulphonate) and poly(vinyl methyl ether-alt- [96] Martens P, Anseth KS. Characterisation of hydrogels formed
maleic anhydride). J Appl Polym Sci 1979;23:2611–25. from acrylate modified poly(vinyl alcohol) macromers. Polymer
[74] Reid CE, Spencer HG. Ultrafiltration of salt solutions by ion- 2000;41:7715–22.
excluding and ion-selective membranes. J Appl Polym Sci [97] Toray Ind Inc., Composite semipermeable membrane–comprises
1960;4(12):354–61. microporous supporting membrane convered thin crosslinking PVA
[75] Immelman E, Sanderson RD, Jacobs EP, van Reenan AJ. Poly(vinyl resin. Jap. Patent JP01,254,203 (1989).
alcohol) gel sub-layers for reverse osmosis membranes. II. Insolu- [98] Trochimczuk AW, Kabay N, Arda M, Streat M. Stabilisation of solvent
bilisation by crosslinking with poly(methyl vinyl ether-alt-maleic impregnated resins (SIRs) by coating with water soluble polymers
anhydride). Desalination 1993;94:37–54. and chemical crosslinking. Reactive Func Polym 2004;59:1–7.
[76] Huang Z, Guan H-M, Tan W-L, Qiao X-Y, Kulprathipanja S. Pervapora- [99] Cheng ATY, Rodriguez F. Mechanical properties of borate crosslinked
tion study of aqueous ethanol solution through zeolite-incorporated poly(vinyl alcohol) gels. J Appl Polym Sci 1981;26:3895–908.
multilayer poly(vinyl alcohol) membranes: effect of zeolites. J [100] Brannon ML, Peppas NA. Solute diffusion in swollen membranes.
Membr Sci 2006;276:260–71. Part VIII. Characterisation of and diffusion in asymmetric mem-
[77] Huang RYM, Yeom CK. Pervaporation separation of aqueous mix- branes. J Membr Sci 1987;32:125–38.
tures using crosslinked poly(vinyl alcohol). II. Permeation of [101] Lebrun L, Da Silva E, Metayer M. Elaboration of ion-exchange
ethanol-water mixtures. J Membr Sci 1990;51:273–92. membranes with semi-interpenetrating polymer networks con-
B. Bolto et al. / Progress in Polymer Science 34 (2009) 969–981 981

taining poly(vinyl alcohol) as polymer matrix. J Appl Polym Sci [110] Myung S-W, Choi I-H, Lee I-C, Kim I-C, Lee K-H. Use of fouling
2002;84:1572–80. resistant nanofiltration and reverse osmosis membranes for dyeing
[102] Kim DH, Na SK, Park JS, Yoon KJ, Ihm DW. Studies on the preparation wastewater effluent treatment. Water Sci Technol 2005;51:159–64.
of hydrolysed starch-g-PAN (HSPAN)/PVA blend films—effect of the [111] Peng P, Fane AG, Xiaodong L. Desalination by membrane distillation
reaction with epichlorohydrin. Eur Polym J 2002;38:119–204. adopting a hydrophilic membrane. Desalination 2005;173:45–54.
[103] Turner JE, Shen M, Lin C-C. Hydrophilic behaviour of HSPAN/PVA [112] Linder C, Perry M, Katraro R. Semipermeable composite mem-
membrane. J Appl Polym Sci 1980;25:1287–96. branes, their manufacture and use. Modified poly(vinyl alcohol)
[104] Hong Y, Shang T, Li Y, Wang L, Wang C, Chen X, et al. Synthesis membranes. US Patent 4,753,725 (1980).
using electrospinning and stabilisation of single layer macroporous [113] Linder C, Kedem O. History of nanofiltration membranes
films and fibrous networks of poly(vinyl alcohol). J Membr Sci 1960 to 1990. In: Schäfer AI, Fane AG, Waite TD, editors.
2006;276:1–7. Nanofiltration—principles and applications. Oxford: Elsevier; 2005.
[105] Gryte CC. 1,1,1-Tris(hydroxymethyl)propane-crosslinked poly(vinyl p. 21–31.
alcohol)–poly(styrene sodium sulphonate) interpolymer mem- [114] Cadotte JE. Alkali resistant hyperfiltration membrane. US Patent
branes. J Appl Polym Sci 1978;22:2401–2. 4,895,661 (1990).
[106] Chao A-C, Yu S-H, Chuang G-S. Using NaCl particles as porogen [115] Kawada I, Inoue K, Kazuse Y, Ito H, Shintani T, Kamiyama Y. New
to prepare a highly adsorbent chitosan membrane. J Membr Sci thin-film composite low pressure reverse osmosis membranes and
2006;280:163–74. spiral wound modules. Desalination 1987;64:387–401.
[107] Weissenborn M, Hutter B, Singh M, Beeskow TC, Anspach FB. [116] Koike S, Subramanian R, Nabetani H, Nakajima M. Separation of oil
A study of combined filtration and adsorption on nylon-based constituents in organic solvents using polymeric membranes. J Am
dye-affinity membranes: separation of recombinant l-alanine dehy- Oil Chem Soc 2002;79:937–42.
drogenase from crude fermentation broth. Biotechnol Appl Biochem [117] Hachisuka H, Ikeda K. Composite reverse osmosis membrane having
1997;25:159–68. a separation layer with polyvinyl alcohol coating and method of
[108] Mansouri J, Fane AG. Osmotic distillation of oily feeds. J Membr Sci reverse osmosis treatment of water using same. US Patent 6,177,011
1999;153:103–20. (1998).
[109] Ikada Y, Iwata H, Mita T, Nagaoka S. Grafting of proteins onto poly- [118] Cuperus FP, Ebert K. Non-aqueous applications of NF. In: Schäfer AI,
mer surfaces with the use of oxidised starch. J Biomed Mater Res Fane AG, Waite TD, editors. Nanofiltration—principles and applica-
1979;13:607–22. tions. Oxford: Elsevier; 2005. p. 521–36.

You might also like