Vous êtes sur la page 1sur 23

Food Eng Rev (2011) 3:136158

DOI 10.1007/s12393-011-9042-8

REVIEW ARTICLE

Fruit Juice Processing and Membrane Technology Application


A. P. Echavarra C. Torras J. Pagan
A. Ibarz

Received: 3 August 2011 / Accepted: 22 September 2011 / Published online: 21 October 2011
Springer Science+Business Media, LLC 2011

Abstract This article provides an overview of recent


developments and the published literature in membrane
technology with regard to fruit juice processing and considers the impact of such technology on product quality. In
the fruit juice industry, membrane technology is used
mainly to clarify the juice by means of ultrafiltration and
microfiltration and to concentrate it by means of nanofiltration and reverse osmosis. We look at enzyme immobilization techniques to improve filtration performance and
operation methods to quantify fouling. Membrane fouling
is a critical issue and inhibits the broader application of
membranes in the fruit production industry. Pectin and its
derivatives form a gel-like structure over the membrane
surface, thereby reducing the permeate flux. In order to
degrade pectin, the raw juice is usually subjected to an
enzymatic treatment with pectinase, which hydrolyses
pectin and causes its protein complexes to flocculate. The
resulting juice has reduced viscosity and a much lower
pectin content, which is advantageous in the subsequent
filtration processes.
Keywords Membrane filtration  Fruit juices  Treatment
enzymes  Fouling  Cleaning

A. P. Echavarra  J. Pagan  A. Ibarz (&)


Food Technology Department, University of Lleida,
XaRTA-UPV, Av. Rovira Roure, 191, 25198 Lleida,
Spain
e-mail: aibarz@tecal.udl.cat
C. Torras
Area de Bioenergia i Biocombustibles, Institut de Recerca en
Energia de Catalunya (IREC), C/Marcel l Domingo,
2, 43007 Tarragona, Spain

123

Introduction
Membrane technology is an alternative to conventional juice
clarification and concentration processes that has been
widely applied in the dairy, and beverage industries since the
discovery of asymmetric membranes by Loeb and Souriragin
in the early 1960s [51]. Membrane processes have been used
in the food industry because they require less manpower, are
more efficient and have a shorter processing time than traditional filtration. Therefore, the operational costs of using
membrane processes are considerably lower than those of
more traditional processes [216]. The types of filtration most
commonly used are ultrafiltration (UF) and microfiltration
(MF), which correspond to pressure-driven processes capable of separating particles in the approximate size ranges of
1100 lm and 0.110 lm, respectively [107]. Wide ranges
of pore size are being used today in the industry, from 18.000
molecular weight cut-off (MWCO) to 0.2 lm [135]. However, if the membrane pore size exceeds 25.000 Da, tannins
can pass into the clarified juice, resulting in a brownish
colour and sharp flavour [107].
Nanofiltration (NF) is a pressure-driven separation process in which the filtration efficiency depends on steric
(sieving) and charge (Donnan) effects. The advantage of
NF over reverse osmosis is that on average it requires 21%
less energy expenditure. As a result, NF is a very promising
process for the food industry and even more so for the
beverage industry, which requires dissolved contaminants
to be reduced and particles to be finely filtered [213]. To
date, NF has been applied for several different purposes in
the juice industry, such as to recover aromas from fruit
juices [57], to treat wastewater from juice beverage production [85] and to regulate sugar concentration [68].
Reverse osmosis (RO) or hyperfiltration is a chemical
engineering unit operation with a wide range of known and

Food Eng Rev (2011) 3:136158

potential commercial applications [12]. It is a separation


technique which operates at (or slightly above or below)
room temperature and can be used to concentrate or purify
liquids without a phase change [99]. Reverse osmosis has
several advantages over the traditional TASTE (Thermally
Accelerated Short Time Evaporator) evaporation technique
[7]. Because less heat is used, the thermal damage to the
products is generally eliminated. The drawbacks of RO are
its inability to reach the concentration of standard products
produced by evaporation because of the limitation of high
osmotic pressure [166, 204]. Unfortunately, the rapid
reduction in permeate flux due to fouling and/or concentration polarization hinders the commercial application of
ultrafiltration in juice processing, and therefore, sufficient
process efficiency for apple juice clarification has not yet
been accomplished [116]. The mechanisms responsible for
membrane fouling are commonly understood as falling into
two general categories: reversible and irreversible fouling
caused by the deposition of colloidal particles, inorganic
and organic compounds, and microbes on the external
surface of the membrane and/or within the membrane pores
[44]. In the apple juice clarification process, membrane
fouling may be caused by pectin, tannins, proteins, starch,
hemicelluloses and cellulose [181]. This process is mainly
limited by the accumulation of matter on the filter,
including concentration polarization and membrane fouling
(formation of a gel layer or a deposit) [33].
To enhance filtration performance, fruit juices are usually treated before filtration with enzyme preparations
aimed at hydrolyzing mainly soluble polysaccharides
responsible for high viscosity [22]. The commercial
enzyme cocktail used for the manufacture of fruit juices
contains three different pectinases: pectinlyase, polygalacturonase and pectinesterase to make pectic complexes
soluble to complete the sedimentation and clarification of
the juice [89].
This review describes the basic concepts of membrane
filtration, highlights some related reported results and
outlines the important factors to consider in the practical
application of membrane processes.

Fruit Juice Components


Fruit juices contain colloids that can lead to problems during
the filtration process. These colloids are either part of the
fruit itself or may be formed by microorganisms during fruit
ripening [17]. Pectin is a structural cell wall polysaccharide
found in all higher plants. It is primarily composed of linear
polymers of D-galacturonic acid linked by a-D-1,4 glycosidic
bonds [218]. Pectins are gel-forming agents present in fruit
juices, whose fibre-like structure impedes the clarification
process by significantly reducing flux and yield [193]. Starch

137

can also lead to difficulties during clarification and can give


rise to a secondary haze in juices and therefore should not be
present in clarified juices [31]. In general, juice yield is
controlled by the concentration factor that can be achieved in
the plant and the characteristics of the juice, such as the pulp
level [215]. The typical concentration of pectin and colloids
present in fruit juice is up to 1% (wt/vol). The type of fruit
will determine the amounts of these constituents the juice
contains [195].
Therefore, juices rich in pectins are depectinized (pretreated with enzymes, pectinase and amylase) before they
are clarified to help reduce viscosity and fouling, resulting
in higher flux [82, 83]. However, the rapid reduction in
permeate flux still occurs, thus hindering the commercial
application of ultrafiltration in juice processing, and an
efficient process for apple juice clarification has not yet
been accomplished [129]. The viscosity of fluid foods is the
result of juice pectin content [176, 186], but in most food
processing units, depectinized juice is used to prevent high
viscosity [172].
Fruit Juice Processing Problems
The browning of fruit is a major problem in the fruit
industry and is believed to be one of the main causes of
quality loss during processing [170]. The mechanism by
which fruit and fruit products turn brown is well understood and can be either enzymatic or non-enzymatic in
origin [52].
Thermal treatments are used in the preservation of fruit
derivatives and in manufacturing operations [96]. The
negative effects of these treatments include non-enzymatic
browning, nutrient loss and the formation of undesirable
by-products such as 5-hydroxymethylfurfural (HMF).
According to the International Federation of Fruit Juice
Producers (IFFJP [101]), an HMF content of 5 mg/L for
juices and 25 mg/kg for concentrates should not be
exceeded [151]. Higher values indicate that the product has
been overheated. In order to minimize deterioration, suitable designs are needed for manufacturing and treatment
equipment [142, 190]. In apple juice concentrate, the
product goes brown during thermal processing due to nonenzymatic browning. During storage, browning is attributed mainly to non-enzymatic reactions [30, 95]. Browning
reactions involve caramelization, ascorbic acid degradation
and the Maillard reaction [2].
Enzymatic Browning (EB)
Enzymatic browning (EB) is the result of fast reactions.
Even optimal processing technology cannot completely
prevent EB during the pulping and pressing of fruit juice,
unless special care is taken to avoid oxygen [127].

123

138

Enzymatic browning starts with the initial enzymatic


oxidation of phenols to quinones by the enzyme polyphenol oxidase in the presence of oxygen [62, 177]. Next,
these quinones undergo further reactions, whether enzymatically catalysed or not, leading to the formation of
pigments of melanins [162, 199]. Polyphenol oxidase
(PPO) is important in the beneficial colouration of some
foods, such as prunes, tea, coffee and cocoa [84, 157]. In
particular, the primary enzyme responsible for the browning reaction is polyphenol oxidase (PPO; EC 1.14.18.1) in
the presence of oxygen. This copper enzyme catalyses the
hydroxylation of monophenols to o-diphenols (cresolase
activity) and the oxidation of o-diphenols to their
corresponding o-quinones (catecholase activity) [30, 45].
Enzymatic browning due to PPO in fruits is controlled in
the food processing industry through the use of ascorbate,
sodium bisulphate and by lowering the pH of the product
through the addition of citric acid [187]. Enzymatic
browning not only has a negative effect on colour, but also
impairs other sensory properties of the product, including
its aroma and texture [50, 75, 111]. Membrane processes
are emerging as significant unit operations for the clarification of fruit juice in the food industry [174]. Depectinization enzymes can be retained by ultrafiltration, as
can polyphenol oxidase (PPO), which causes enzymatic
browning [214, 220]. Cloudy apple juice has gained an
increasing market value due to its sensory and nutritional
qualities [162]. Although a typical amber-like hue is
commercially desirable in clarified apple juice, both apple
puree and cloudy juice are expected to have the yellowish
colour which characterizes the fresh product [128].
The control of enzymatic browning is of great importance at the start of these processes. One approach for the
prevention of the enzymatic browning of fruit juices has
been the use of anti-browning agents [40, 162]. Extensive
research [143, 168] has focused on controlling browning by
targeting PPO, substrates (oxygen and phenols) and the end
products of the browning reaction. Based on the working
mechanisms, browning inhibitors can be categorized into
six groups: reducing agents, acidulants, chelating agents,
complexing agents, enzyme treatments and enzyme inhibitors [82]. Of all the inhibitors tested, reducing agents (e.g.
ascorbic acid and its derivatives, cysteine and glutathione)
have been found to be most effective in controlling
browning, and a calcium-ascorbate-based formula has been
widely used by the fresh-cut apple industry [130]. Some
authors have studied the use of ultrafiltration as an alternative to sulphiting for the control of enzymatic browning
[72, 146]. UF is believed to remove polyphenol oxidase
and to be effective in stabilizing the colour of fruit juices
[5, 26, 154], but not with lower molecular weight polyphenols or Maillard reaction precursors, which can undergo
non-enzymatic browning during the storage of the fruit

123

Food Eng Rev (2011) 3:136158

juice [127]. Galeazzi et al. [67] found that banana PPO


fractions had molecular weights [30 kDa, within the range
of molecular weight cut-offs for UF membranes.
Non-Enzymatic Browning (NEB)
During the manufacturing process, changes take place in
the structure of derivative fruit products, and thermal
treatments during preservation processes can affect the
quality of juices and fruit purees through non-enzymatic
browning reactions [141, 170]. NEB in fruit products can
be inhibited or reduced by refrigeration and by the control
of water activity in dehydrated fruits [127]. Two basic
treatments have been used for the control or reduction of
non-enzymatic reactions [127]: the reduction of amino
nitrogen in juices and the use of different chemical inhibitors [69, 79].
The Maillard reaction, which takes place between amino
groups and reducing sugars, is the primary cause of
browning in apple juice [2, 30]. This reaction improves the
desirable sensory characteristics of these foods, such as
colour, aroma and flavour, but it can also have undesirable
side effects, as the Maillard reaction gives rise to losses in
the nutritional value of foods [140]. Studies have also
revealed that Maillard reaction products have antimicrobial
and antioxidant properties [64]. Since quality is very
important in food products, deterioration must be controlled during storage [69, 81].
The search for effective and safe inhibitors of NEB is
driven by the need to: (1) prevent undesirable Maillard
reactions and (2) find alternatives to the use of sulphites
[169]. Organoleptic considerations are the main barrier to
the use of some Maillard inhibitors [127]. Cysteine appears
to be a good alternative to the use of sulphites in foods at
present [191].

Techniques for Improving Fruit Juice Flux


During Membrane Clarification
Enzymes Used in Fruit Processing
The widespread and more efficient use of microbial
enzymes in the preparation of wines and fruit juices started
in 1960s. As a result, pectinases are today one of the
leading enzymes used in the commercial fruit juice processing sector (Fig. 1) [106]. These enzymes are primarily
responsible for the degradation of the long and complex
molecules called pectins, which occur as structural polysaccharides in the middle lamella and in the primary walls
of young plant cells [165]. It has been suggested that by
using pectinase enzymes, the pectin-coating surrounding
the protein particles is broken down, as shown in Fig. 2.

Food Eng Rev (2011) 3:136158

139

Fig. 1 General fruit processing


line steps (modified from
Lozano [127])

This process allows the particles to aggregate and accumulate as sediment in the bottom of the tank [16]. Pectins
can cause problems in the food industry by giving rise to
turbidity and viscosity during the extraction, filtration and
clarification of fruit juices [65]. In fact, the presence of
pectic substances in fruit juices causes a considerable
increase in their viscosity, thereby impeding filtration and
subsequent concentration processes [3, 21]. Treatment of
the mashed fruit with macerating enzymes further breaks
down the fruit pulp, resulting in increased juice yields,
reduced viscosity and improved run-off [182]. The dose,
time and temperature should be determined in trials
depending on the degree of maceration required and the
type of fruit used [8, 14]. Liquefaction can be achieved,
although the enzymes are equally beneficial to yield
improvement for juice extraction and viscosity reduction
[182]. In contrast, the development of turbidity during cold
storage, usually referred to as haze formation, is assumed
to be caused by interactions between haze-active proteins
and polyphenols that form insoluble multi-molecular
structures [103].
Clarification is affected by pH, temperature, contact
time and enzyme concentration. A juice with a low pH will
be clarified more readily than one with a higher pH. The

optimal activity of these enzymes is at pH 45 (the natural


pH of the fruit) and at temperatures below 50 C [39, 106].
Application of Pectinases
Acidic pectic enzymes used in the fruit juice industry and
in winemaking often come from fungal sources, particularly from Aspergillus niger [185]. The main juices commercially produced by these industries include sparkling
clear apple, pear and grape juices. Enzymes are added in
order to increase the juice yield during pressing and
straining and to remove suspended matter to provide
sparkling clear juices (free of haze) as well as cloudy juices
(citrus, prune and tomato juices and nectars) [106].
Pectinases are used in the clarification of fruit juice
before membrane filtration by adding them to the bulk juice
with a minimum amount of mechanical agitation [20]. Fruit
cell walls have such a complex structure of interwoven
polymers that no single enzyme will break down cell walls
[29]; therefore, the complete liquefaction of a cell wall is
likely to require a complex cocktail of carbohydrase
enzymes [145]. The classification of pectic enzymes is
based on their attack on the galacturonic backbone of the
pectic substance molecule (Table 1). Basically, three types

123

140

Food Eng Rev (2011) 3:136158

Fig. 2 Theory of floc formation


during enzyme treatment of
juice (adapted from Barrett et al.
[16])

Table 1 Enzymes used in fruit processing


Enzymes

Characteristics

Polygalacturonase
(PG)

Responsible for the random hydrolysis of alpha-1,4-glycosidic linkages between galacturonic acid residues.
Depolymerize low esterified pectin (endo- and exo-enzymes)

Pectin lyase (PL)

Cleaves the pectin, by an elimination reaction releasing oligosaccharides with non-reducing terminal alpha-1,4-linked
galacturonic acid residues, without the necessity of pectin methyl esterase action

Pectin methylesterase
(PE)

Releases methanol from the pectyl methyl esters, a necessary stage before the polygalacturonase can act fully (the
increase in the methanol content of such treated juice is generally less than the natural concentrations and poses no
health risk)

Xylanase
(hemicellulase)

A mixture of hydrolytic enzymes including xylan endo-1,3-beta-xylosidase and xylan 1,4-b-xylosidase, which degrade
hemicellulose

Arabonases (ARA)

Hydrolyse arabinans

Ferulic acid esterase


(FAE)

Cuts ferulic acid and other phenolic linkages between the xylan chains opening the structure to further degradation by
xylanases.

Cellulase

Breaks down cellulose

Amylases

Breaks down starch

of pectic enzymes exist: de-esterifying enzymes (pectinesterases), which catalyse the de-esterification of the
methoxyl group of pectin forming pectic acid and which
are produced by fungi, bacteria, yeast and higher plants;
depolymerizing enzymes (pectinases: hydrolases and lyases); and protopectinases [90].
Commercial pectinases are usually mixtures of
enzymes, including pectinase, polygalacturonase, pectinesterase, pectin lyase, cellulase, protease and amylase,
which cause suspended components in the juice to
agglomerate, form a floc and settle out. Furthermore, the
recovery of this group of enzymes from the membrane
clarification process has been shown to increase pressing
efficiency and reduce industrial processing costs [90, 163].
Polygalacturonase treatment has been found to decrease
the size of granule particles in Spartan apple juice as well
as to remove their web-like aspect. The reduction in the

123

size of the particles and the subsequent decrease in viscosity could partially explain how depectinization
improves flux [70]. Maceration is a process in which
organized tissue is transformed into a suspension of intact
cells, resulting in pulpy products used as a base material for
pulpy juices and nectars [148]. Enzymes are also used to
help clarify and stabilize the juice by degrading soluble
pectins and starches that would otherwise cause haze [55].
The reported improvements in juice flux by means
of maceration and depectinization are summarized in
Table 2. Landbo et al. [119] pre-press macerated black
currant juice with ten different pectinolytic enzymes.
The optimal maceration was achieved using an enzyme
dosage of 0.18% by wet weight of berries with a reaction at 60 C for 30 min on the most finely crushed
berry mash. Juice yields ranged from 66.4 to 78.9% by
wet weight of mash. Depectinization is therefore needed

Food Eng Rev (2011) 3:136158

141

Table 2 Effect of enzymatic pre-treatment in fruit juice clarification


Pre-treatment

Comments

References

Maceration

Black currant juice was treated with pre-press maceration with ten different pectinolytic
enzymes. The cloned Aspergillus niger/Aspergillus aculeatus preparation Pectinex BE
(1,000 FDU 55 C/mL activity) consistently resulted in the best responses regarding
anthocyanin yields, phenols and low juice turbidity. Pectic enzymes containing high levels
of polygalacturonase activity are added to fruit juices to stabilize the cloudiness of citrus
juices, purees and nectars.

Landbo et al.
[119]

Depectinization

Cherry juice was treated with Pectinex Smash, a protease preparation derived from
Aspergillus spp. The effects of the alternative clarification treatments were assessed
immediately after the particular clarification treatment (immediate turbidity). The protease
treatment resulted in a significant reduction in immediate turbidity. Protease addition may
be a workable alternative measure for decreasing immediate turbidity levels.

Pinelo et al. [167]

Depectinization

The influence of enzymatic treatment using enzyme concentrations of 20, 100 and 300 mg/L,
a time of 90 min and a temperature of 40 C for depectinization was studied in cherry juice
and pineapple (Ananas comosus) juice. A lower permeate flow rate was found for the
polysulphone hollow fibre membrane. The increase in permeate flow rate with the use of
the 300 and 100 mg/L enzyme concentration was not significant, so it is economically
advantageous to ultrafilter depectinized juice treated with an enzyme concentration of
20 mg/L.

Barros et al. [17]

Depectinization

Peach juice was clarified using the enzymatic hydrolysis Pectinex AFP L3 (Novozymes) to
reduce the viscosity of the juice and its pulp content, and consequently to increase juice
extraction. Physical and chemical analyses showed that enzymatic treatment is effective for
reducing peach pulp viscosity, pulp content and turbidity and does not influence other juice
parameters such as pH, total acidity, vitamin C and soluble solids. Hydrolysis using at
25 C for 60 min yielded the best results for the reduction in pulp (48%) and viscosity
(68%).

Santon et al. [181]

Depectinization

To prevent fouling the membrane, black currant juice was depectinized with Panzym Super E
liquid enzyme preparation. The ultrafiltration was carried out at a transmembrane pressure
of 2 bars and an operating temperature of 25 C. The effect of processing on the valuable
anthocyanin and flavonol content of the juices was examined. The results indicate that
enzymatic treatment increases the valuable compound content of the juice. However, the
ultrafiltration process resulted in the significant loss of other valuable content. 54% of total
flavonol and 50% of total anthocyanins were maintained in ultrafiltered juice when
compared with the feed samples.

Pap et al. [164]

Enzyme immobilization

The catalytic behaviour of a mixture of pectic enzymes in a pectin aqueous solution


covalently immobilized on different supports was analysed. No differences were observed
at low circulation rates, while at higher recirculation rates, the time required to obtain
complete pectin hydrolysis in the fluidized reactor was found to be 0.25 times smaller than
in the packed bed reactor: 131 min for the packed reactors and 41 min for the fluidized
reactors.

Diano et al. [56]

Co-immobilization

The process was applied in the clarification of apple juice and the potential of coimmobilized pectinase amylase by physical adsorption on a Uf hollow fibre membrane was
studied. The concentration of reaction products increased by up to 50% with the pectin
concentration, and the starch content changed from 3.85 to 5.00 mg/mL. The reference
permeate flux was improved when starch was added to the substrate, regardless of its
concentration.
Apple juice and model solution containing pectins were assessed through tubular membranes
of 100 and 300 kDa (one and three channels). In order to obtain a higher permeate flow, the
fluids were pre-treated with pectinolytic enzymes recirculating through a tubular membrane
at different concentrations and treatment times. Depectinization increased the permeate
flow for the model solution and apple juice with 67.52 and 53.11% when the pectinolytic
enzyme preparation re-circulated across the tubular membrane.

Carrn et al. [32]

Previous enzymatic
recirculation

to achieve high flux and concentration factors in membrane clarification processes. Different kinds of fruit
juices have been studied: cherry juice [167], cherry and
pineapple juices [17], peach juice [181] and black currant juice [164]. Permeate flux can be enhanced by pretreating the feed. This technique is commonly used

Echavarria et al.
[61].

either to remove particles that may cause clogging in the


module or to prevent particles or macromolecules from
reaching and being deposited on the membrane modules.
Some authors have described enzyme immobilization
[56], co-immobilization [32] and prior enzymatic recirculation [61].

123

142

Food Eng Rev (2011) 3:136158

The use of innovative pre-treatments can significantly


extend the UF volumes that can be inexpensively treated
[25]. Furthermore, further research into enzymatic membrane reactors with activated carbon is needed with a
membrane that holds an active enzyme, either by light or
by strong bonding. In addition to the intrinsic advantages of
these systems, the process is continuous, the catalyst
component can be re-used and a permeate free of this
compound is obtained [201].

Applications of Membrane Processes for Fruit Juice


Processing
Membranes
The most important parameters that characterize a membrane are its permeability and its separation capability
[153].
The measurable parameter for flow is volumetric or
weight flow rate (L/h or g/L). However, in filtration
experiments, where pressure is the driving force, this value
cannot be compared with other processes because it
depends on the membrane area as well as transmembrane
pressure (TMP). To normalize the flow rate, the flux term
[70] (flow rate normalized by membrane area, L/hm2) and
particularly the permeability term (flux normalized by
TMP, L/hm2 at 1 bar) are used.
Transmembrane pressure (TMP) is one of the factors
that influences permeate flux during UF and MF. Permeate
flux increases with TMP, but it also decreases with the
increasing resistance of the membrane [122].
Permeate flux increases with the TMP, but the relation
between them is only linear when the feed is pure water
[125]. At the higher pressure, the flux becomes independent
of the pressure the system being in the mass-transfer-controlled region, where the concentration polarisation layer
reaches a limiting concentration [180] (Fig. 3).
Membrane separation capability is normally established
by using the molecular weight cut-off (MWCO) concept,
which corresponds to the size of particles that are rejected
by the membrane 90% of the time [175]. To determine the
MWCO of a membrane, a mixture of several compounds
with different and narrow molecular weights (i.e. polyethylene glycols) are tested with the membrane in question.
The use of MWCO to measure membrane separation
capability is suitable with membranes with low pore sizes
[210], but for membranes with larger pore sizes, MWCO
often does not properly describe the separation capability
of a membrane because other parameters, like the type of
solvent, have such a strong influence. In those cases, the
mean pore size of the membrane is used directly [49, 123].
The porous membrane is the filtration medium, the fluid

123

Fig. 3 Relationship flux and parameters during membrane filtration,


indicating the areas of pressure control and mass transfer control
(modified from Mulder [150])

that crosses the porous membrane and does not contain


retained solids is the filtrate or permeate [135, 196], while
the remaining fluid corresponds to the retentate or concentrate that can be recirculated to extract as much permeate as possible [211]. Generally, the pore size of MF
membranes ranges from 0.1 to 10 lm [117]. UF membranes have smaller pores than MF membranes, ranging
from 0.1 to 0.01 lm. The typical range of MWCO levels is
less than 100 Daltons for RO membranes, and between 200
and 1,000 for NF membranes [222].
Additionally, the coefficient of retention (R) can be
calculated as (Eq. 1) [93]:
R1

Cp
Cc

where Cp is the concentration of the component of interest


(mass solute/volume dissolution) and Cc (mg/L) is the
concentration in the retentate.
High flux and product quality must be maintained with
UF or MF to be considered an effective clarification
method [25]. As membrane pore size increases, flux
improves due to lower flow resistance [218]. If particles are
present, flux may not increase proportionally with pore size
due to membrane fouling [28].
Many researchers have suggested the optimal operating
parameters (MWCO membrane, temperature, TMP, crossflow velocity) for the membrane clarification of different
kinds of fruit juices (Table 3).
Nong-xue et al. [156] studied the separation conditions
of ultrafiltration for apple pectin, and the relationships
between apple pectin and different relative molecular
weights and their structures and properties. Galacturonic
acid content and degrees of esterification and gelatination
increase as the molecular weight of apple pectin increases,
and the relative monosaccharide composition increases
much more, but the galacturonic acid content in pectins
with a molecular weight of 510 kDa was only lower than
that in pectins with molecular weights of 100300 kDa and

Food Eng Rev (2011) 3:136158

143

Table 3 Optimal operating parameters reported during membrane clarification of fruit juice
Juice

Membrane (kDa)

Optimal parameters
a

TMP

CFV

T
( C)

Comments

References

Apple
pectin

51030100300

0.02, 0.04,
0.06, 0.08
and 0.1 MPa

27 m/s

50

Galacturonic acid content in pectins with molecular


weight of 510 kDa was only lower than that in
pectins with molecular weights of 100300 kDa
and over 300 kDa. Apple pectins with molecular
weights over 300 kDa were in the majority,
accounting for 68.04%, and those with molecular
weights below 5 kDa were the minority, only
accounting for 3.31%.

Nong-xue
et al. [156]

Apple
golden
delicious

10100 (PES)b

nde

nde

50

With 10 kDa membrane, colour was improved and


some minor components, mainly phenolics and
trans-2-hexenal, were retained.

Onsekizoglu
et al. [161]

The composition of the juice was only slightly


changed after UF with 100 kDa.
Kiwifruit

15 (PVDF)c tubular
ceramic

0.85 bar

800 L/h

25

The results showed that flux increased with


temperatures from 20 to 30 C and with axial feed
flow rates from 300 to 700 L/h.

Cassano
et al. [35]

Apple

1550 Zirconium
dioxide (ZrO2)

150400 kPa

27 m/s

nde

The average permeate flux varied between 56 and


157 L/(m2h). The results indicate the condition for
the set of variables (colour, clarity and turbidity of
juice) improved significantly after UF.

De Bruijn
et al. [54]

Pineapple

50100 (PS)d

1.53.0 bar

nde

25

Sugar loss was evaluated (glucose, fructose and


sucrose) in clarified pineapple juice. Sugar content
was found to decrease most when the 3080-kDa
tubular membrane at 1.5 bar was used for
clarification. The best total sugar recoveries have
been observed in juices clarified with polysulfone
membranes (50 kDa7.5 bar).

Carvalho
et al. [34]

350 kPa

13.5 L/h

50

Depectinization increased the permeate flow for the


model solution and apple juice with 67.52 and
53.11%. Using the 300-kDa membrane with three
channels increased the density and flow rate of the
permeate at the lower treatment time. In the
1-channel 300- and 100-kDa membrane, there were
no appreciable colour changes.

Echavarria
et al. [61]

3080 (PESPVDF)

Apple and

100300

Model
solution

(1-channel)
300 (3-channels)
Zirconium dioxide
(ZrO2)

Cross-flow velocity

Polyethersulphone membrane

Polyvinylidenefluoride membrane
Polysulfone membrane

d
e

Undetermined

over 300 kDa. The galacturonic acid content in pectins


with molecular weights of 30100 kDa, 1030 kDa, and
below or equal to 5 kDa was likely to be polysaccharose.
Most apple pectins had molecular weights over 300 kDa,
accounting for 68.04%, and those with molecular weights
below 5 kDa were in the minority, only accounting for
3.31%.
12 Brix apple juice was ultrafiltrated by Onsekizoglu
et al. [161] through 10-kDa and 100-kDa MWCO membranes. Physical quality attributes such as colour were
especially improved through the 10-kDa membrane. Only
some minor components, mainly phenolics and trans-2hexenal, were retained by the same membrane. The

composition of the apple juice was barely changed after UF


with 100 kDa. Cassano et al. [35] clarified fresh depectinized kiwifruit juice using a UF process with a 15-kDa
MWCO membrane and found that flux increased with
temperatures from 20 to 30 C and 90 kPa of pressure with
an axial feed flow rate of from 300 to 700 L/h. De Bruijn
et al. [54] used a 15- and a 50-kDa MWCO membrane to
clarify apple juice with a tangential velocity of 2 and 7 m/s
and a TMP of 150 and 400 kPa. The average permeate flux
varied between 56 and 157 L/(m2h), and a set of variables
(colour, clarity and turbidity of juice) was found to have
improved significantly after UF. Jaeger de Carvalho et al.
[34] evaluated the loss of sugars (glucose, fructose and

123

144

sucrose) in pineapple juice (Ananas comosus, L. Merril)


after clarification by cross-flow micro- and ultrafiltration
using two different module geometries (plate and frame
and tubular systems) to select the membrane process that
would best preserve these nutrients. The membrane pore
diameters and cut-offs were 0.1, 0.45 lm, and 50, 100 kDa
(PS), and 0.3 lm and 3080 kDa (PES and PVDF). All
processes were operated at different TMP and at room
temperature (25 2 C). It was observed that the sugar
content was more reduced when the 3080-kDa tubular
membrane at 1.5 bar was used for pineapple juice clarification. Although the best total sugar recoveries have been
observed in juices clarified with polysulfone membranes
(50 kDa7.5 bar), the use of a 0.3-lm PES, due to its
tubular configuration and module geometry, is more
attractive and appropriate. A model solution and apple
juice containing pectins were ultrafiltrated by tubular
membranes of 100 and 300 kDa (one and three channels)
to obtain a higher permeate flow [61].
Depectinization of the model solution and apple juice
increased the permeate flow by 67.52 and 53.11% when the
pectinolytic enzyme preparation re-circulated across the
tubular membrane. Using the 300-kDa membrane with
three channels increased the density and flow rate of the
permeate at the shorter treatment time. In the 1-channel
300-kDa membrane, unlike the 1-channel 100-kDa membrane, there were no appreciable colour changes.
Separation Processes by Membranes
The filtration techniques reviewed here include not only the
familiar solid/liquid separation processes that form the
basis of filtration, but also the closely related diffusion
processes of reverse osmosis, nanofiltration, microfiltration
and ultrafiltration with membranes [209]. In general, three
filtration modes can be distinguished: unstirred and stirred
dead-end filtration and cross-flow filtration [60]. In
Fig. 4 a Dead-end and crossflow filtration and b cross-flow
filtration (modified from Saxena
et al. [183])

123

Food Eng Rev (2011) 3:136158

unstirred dead-end filtration, the test fluid is put under


pressure without any agitation of the liquid [160]. In the
stirred dead-end mode, the fluid is agitated with a stirring
bar [1]. In dead-end (through-flow) filtration, the feed
stream moves perpendicularly to the barrier (Fig. 4a). The
build-up of a cake layer on the membrane barrier will
restrict the filtration rate or flux [58]. In cross-flow filtration, the solution is pumped to flow tangentially over the
membrane surface (Fig. 4b). Separation processes based on
membrane technology, especially cross-flow ultrafiltration,
have proved suitable for the separation of delicate materials
and are widely applied at an industrial scale due to their
easy scalability, economic advantageousness and nondenaturing behaviour [139]. Dead-end filtration is still
quite common for MF, but cross-flow MF systems are also
employed [120]. According to Girard et al. [70], Daufin
et al. [53] and Schafer et al. [184], a combination of
membrane operations is commonly applied (successive
filtrations with different MWCO membranes or with other
types of separation: pre-treatment by enzyme, depectinization, concentration by evaporation, ion exchange for
demineralization, de-acidification, discolouring, etc.) in the
fruit, vegetable and sugar juice industries. Depending on
the size of the pores in a membrane, different components
can be fractionated and/or concentrated. MF and UF have
the ability to remove suspended matter and colloids from
the feed stream in water treatment systems via a sieving
mechanism based on the size of the membrane pores [102].
The primary applications of membrane operations in the
food industry are in the beverage industry (wine, beer, fruit
juices, etc.), the dairy industry (whey protein concentration, milk protein standardization, etc.) and, to a lesser
extent, in the processing of egg products. The use of UF or
MF as an alternative to clarified fruit, vegetable and sugar
juices greatly simplifies the operational process and results
in increased juice yield and improved product quality [121,
221]. Enzymatic hydrolysis combined with selective UF

Food Eng Rev (2011) 3:136158

145

Table 4 Classification of membrane separation processes (adapted and modified from Mulder [150])
Characteristics

Microfiltration

Ultrafiltration

Nanofiltration

Reverse osmosis

Membrane

Symmetrical
asymmetric

Asymmetrical

Asymmetrical

Asymmetricacomposite

Thickness

Sa & 10150 lm

S & 150250 lm; Tb & 1 lm

S & 150 lm;


T & 1 lm

S & 150 lm; T & 1 lm

Pore size
Driving force
(pressure)

0.110 lm
\2 bar

0.010.1 lm
110 bar

\0.001 lm
540 bar

\0.001 lm
10100 bar

Rejects

Particles, clay,
bacteria

Macromolecules, proteins, polysaccharides,


viruses

HMWC mono-di and


oligosaccharides,
polyvalent anions

HMWCc, LMWCd,
sodium, chloride,
glucose, amino acids,
proteins.

Membrane
material

Ceramic, PPe,
PS(PSO)f, PVDFg

Polymer (e.g. polysulfone, polyacrylonitrile),


PVDF, PS(PSO), PESh, ceramic (e.g.
Zirconium oxide, aluminium oxide)

Cellulose triacetate,
aromatic polyamide,
polyamide

Cellulose triacetate,
aromatic polyamide,
polyamide

Thin film

Thin film

Membrane
module

Tubular, hollowfibre, spiral,


wound, plate and
frame

Tubular, hollow-fibre, spiral-wound, plate and


frame

Tubular, spiral-wound,
plate and frame

Tubular, spiral-wound,
plate and frame

Main
applications

Analytical
applications

Dairy (milk, whey, cheese making)

Concentration fruit
juice

Concentration dairy
products (milk)

Sterilisation (foods
and
pharmaceuticals)

Clarification (fruit, vegetables and oils)

Demineralisation of
water sugar solutions
and whey

Concentration of food juice


and sugars

Clarification
(beverages)

Food (potato starch and proteins)


Pharmaceutical (enzymes, antibiotics,
pyrogens)
Water treatment

Cell harvesting
(biotechnology)

Recycle of nutrients in
fermentation
processes

Desalination of brackish
and seawater
Production of ultrapure
water (power industry)

Separation of
sunflower oil from
solvent
Treatment of
wastewater

Sublayer

Top layer

High molecular weight compound: 100,0001 million mol/g

Low-molecular weight compounds: 1,000100,000 mol/g, macromolecules: 1 million mol/g

Polypropylene membrane

Polysulfone (either polyethersulfone or polyarylethersulfone)

Polyvinylidenefluoride membrane

Polyethersulphone membrane

can produce beverages from vegetable proteins. In the beer


industry, recovery of maturation and fermentation tank
bottoms is already applied at industrial scale [53]. Applications of NF include wastewater treatment and lactic acid
recovery from fermentation media to clarify and concentrate as well as to decolourize juices [27]. A summary of
membrane processes is shown in Table 4, adapted from
Mulder [150]. Many researchers [77, 121] have suggested
that the use of membrane separation processes to clarify
and concentrate liquids is less costly than conventional

filtration and vacuum evaporation, justifying their application in the food industry.
Ultrafiltration (UF) and Microfiltration (MF)
UF and MF with tangential filtration or cross-flow
filtration have been replacing conventional filtration
methods for clarifying fruit juice [171]. An effective
clarification method must maintain high flux and product
quality [59]. The advantages of UF and MF over

123

146

conventional methods include a reduction in enzyme


consumption, the elimination of fining agents and their
associated problems, and production with a continuous
simplified process [161]. Microfiltration can be used
to separate juices into a fibrous concentrated pulp and
a clarified fraction free of spoilage microorganisms,
improving the microbiological quality of the permeate
juice [34]. Microporous membranes of different materials,
configurations and MWCOs can be used. The feeding
solution is applied in parallel to the membrane surface,
and pressure is the primary driving force [18, 33]. Some
changes can be expected to take place in the juices when
either method is applied [147]. Hakimzadeh et al. [80]
purified raw sugar beet juice using a ceramic membrane in
a MF process and a polymeric membrane in a UF process.
According to their reports, MF reduced the total soluble
solids and turbidity and altered the colour of the juice
more than the UF process. Cassano et al. [38] clarified
cactus pear (Opuntia ficus) using both MF and UF processes. That study looked into the effect of microfiltration
(MF) and ultrafiltration (UF) processes on the physicochemical composition of the juice and the influence of the
clarification treatment on the content of the main parameters of the nutritional and functional properties of the
fruit. The permeate of both processes was a clear solution
with a low protein content enriched with antioxidant
compounds (polyphenols, vitamin C, etc.), sugars, amino
acids and minerals. However, Mirsaeedghazi et al. [147]
used the above techniques to process pomegranate juice,
which has valuable components (such as antioxidants) that
can be removed during membrane processing. According
to the results of this study, there was a greater membrane
fouling in the UF process when compared to the MF
process.
Pre-treatment of the juices with an MF process reduced
the membrane fouling in the UF process. However, further
processing of the microfiltered pomegranate juice by a UF
process did not have any influence on the turbidity but
decreased the levels of valuable components (such as total
anthocyanin and antioxidant activity levels) of the juice.
Furthermore, different membrane processes (MF or UF)
can result in different efficiency levels in reducing juice
clarity and in membrane fouling.
Wang et al. [212] clarified West Indian cherry juice
using a 0.14 lm cut-off tubular ceramic MF membrane.
The optimum and highest permeate flux by MF is achieved
at 40 C. Following clarification with MF, West Indian
cherry juice retains a high concentration of the ascorbic
acid and becomes an attractive light pink colour. Both
estimated and measured component concentrations in the
permeate and retentate display good fit; thus, this mathematical model can be theoretically applied in the MF
system for selecting operating parameters.

123

Food Eng Rev (2011) 3:136158

Ultrafiltration has also been investigated for the clarification of orange and lemon juice [48, 66], pineapple [11],
clementine mandarin (Citrus reticulata) [37], kiwifruit
[198] and apple [61, 161].
Nanofiltration (NF) and Reverse Osmosis (RO)
Nanofiltration (NF) and reverse osmosis (RO) are known as
hyper-filtration and are often applied to remove dissolved
contaminants from a feed stream, as in the case of softening or desalination [109]. NF is a process used to concentrate juices, and in the sugar industry, NF has been used
as one of the steps in the clarification and concentration of
raw juice as well as in the processing of non-sugar compounds [78]. NF features quite distinctive properties such
as pore radius and surface charge density which influence
the separation of various solutes [91, 100]. The separation
of solutes in the NF range is dependent upon the microhydrodynamics and interfacial events occurring on the
membrane surface and inside the membrane pore [188]. A
study on the possible application of PA300 and FT30
reverse osmosis membranes in the sugar industry found
that this process concentrates sugar juice to about 1330%
sugar content [133]. Other studies have found that sugar
losses can be reduced, and the size of the evaporators
minimized to recover aromas from fruit juices [15] and that
wastewater from beverage production can be treated using
the same processes [27].
RO is a pressure-driven membrane process that can be
used as an alternative process for juice concentration, as it
does not involve phase change or the use of high temperatures [98]. The major components in fruit juices contributing to osmotic pressure are sugars (hexoses and dissaccharides) and organic acids. The main advantages of
RO concentration are the attainment of high-quality products due to low operating temperatures, resulting in the
retention of nutritional, aroma and flavour compounds,
lower energy consumption and the use of easily operated,
compact facilities [6]. On the other hand, a major disadvantage of RO is the lower concentration level that it
obtains compared to thermal evaporation [98]. In conventional juice processing, the concentration levels of the fruit
juice range from 42 to 65 Brix. Therefore, RO should be
viewed as a first-stage process combined with other processes like osmotic evaporation [70]. RO requires high
operating pressures in order to overcome the osmotic
pressure of the juice concentrate (ranging from 10 to
200 bar) [144]. Gurak et al. [77] studied the concentration
of grape juice by reverse osmosis with TMPs of 40, 50 and
60 bar and temperatures of 20, 30 and 40 C. Their results
showed that the best processing conditions were at 60 bar
and 40 C, based on the resulting high permeate flux value
and the grape juice concentration of up to 28.5 Brix.

Food Eng Rev (2011) 3:136158

147

Under these conditions, an increase in permeate flux was


achieved with no significant difference in the physical or
chemical parameters of the product compared to the best
conditions found in the factorial design. High-quality
concentrated fruit juices with soluble solids contents
ranging from 28 to 35 Brix were obtained by Sa et al.
[179] and Kozak et al. [113].
The rejection rate for RO is often measured using NaCl
and can reach 99.7% or more. NF replacement or combining this process with RO decreases energy consumption,
reduces sugar losses and makes it possible to minimize the
size of the evaporators during raw juice clarification and
concentration.
Limiting flux phenomena can restrict its application
[208]. Limiting flux phenomena are caused by membrane
properties, solute characteristics and process parameters.
The effects of these three factors on the membrane separation process have been discussed in detail [138]. The flux
of the solvent (JD) in (L/hm2) through RO would be Eq. 2
[93]:
JD ADP  Dp

where A is the membrane permeability coefficient, (DP) is


the pressure drop through the bed, and Dp is the osmotic
pressure.
The solute flux (JS) is determined by Eq. 3 [93].
JS BCc  Cp

where B is the membrane permeability to the solute, Cc is


the concentration in the solution and Cp is the concentration in the permeate.
A concentration gradient of these solutes is formed with
the passage of the solvent through the membrane as illustrated by Fig. 5. The gelled polarized layer offers a new
resistance to mass transfer. Where (ks) is a global transfer
coefficient and (Cg) is the rejected compound in the layer.
Where the transfer phenomenon through the gel layer

Fig. 5 A schematic representation of concentration polarization.


From Ibarz and Barbosa-Canovas [94]

controls the global process, the flux will be expressed as


Eq. 4:
 
Cg
4
J kS ln
Cc
In order to avoid the build-up of concentration polarization layers, the operation should be done at high feeding
velocities [95].

Operating Parameters of Membrane Processes


Optimizing operation conditions may involve limiting
transmembrane pressure, controlling temperature and maintaining a high cross-flow velocity by minimizing fouling
and concentration polarization. This can be achieved by
corrective methods like periodic hydraulic and/or mechanical cleaning (including the choice of cleaning chemicals and
frequency) [223] or by preventive methods like an optimum
design of the membrane module through a fluid dynamics
analysis or dynamic filtration [100]. Concentration polarization and cake formation in the membrane module is controlled either by increased shear at the membrane surface, by
eliminating stagnant points or by using turbulence inducers.
Increased shear can be achieved by pumping the feed at
higher flow rates or by using thin flow channels above the
membrane surface [203]. Many authors propose a prevention
method to maximize the effective membrane area but also to
prevent the decrease in flux during the use. One prevention
method consists of using a suitably designed membrane
module to ensure suitable flow over the membrane [201].
Rotating disc filtration modules have been shown to yield
higher permeate fluxes and better solute transmission than
conventional cross-flow filtration due to the high shear rates
they generate on the membrane which prevent or limit cake
formation in microfiltration [183, 201]. In addition, these
shear rates are obtained at relatively low inlet flow rates and
low pressure drops inside the module, resulting in a relatively
uniform transmembrane pressure (TMP). The combination
of high shear rates and low TMP facilitates macromolecule
transmission through the membrane [97, 200].
The performance of membrane processes can also be
improved when fluid instabilities are superimposed on
cross-flow [211]. In cross-flow membranes, transport
depends on two major actions: moving the particles
towards the membrane surface (negative direction) and
shifting them away from the membrane surface (positive
direction). The negative direction forces include van der
Waals attraction and permeation drag, while the positive
direction actions include electrical double layer repulsion,
Brownian diffusion, shear-induced diffusion and lateral
inertial lift [152].

123

148

Feed/retentate temperature was studied by Goosen et al.


[73], who showed that polymer membranes can be very
sensitive to changes in feed temperature. An increase of up
to 60% in the permeate flux was found when the feed
temperature was increased from 20 to 40 C. This occurred
both in the presence and in the absence of solute. Permeate
flux appears to reach its minimum rate at an intermediate
temperature. Doubling the feed flow rate increases permeate
flux by up to 10%, but only at a high solute concentration.
The high feed/retentate velocities used in UF modules
require high pressure, but low transmembrane pressure is
needed to prevent the compaction of deposits on the
membrane.
De Bruijn et al. [54] reported the effects of cross-flow
velocity, transmembrane pressure and molecular weight cutoff (MWCO) on membrane fouling in the UF of apple juice.
The results showed that membrane fouling was low at high
feed velocity (7 m/s) and low transmembrane pressure
(150 kPa). The optimal conditions of juice permeation were
400 kPa and 2 m/s. At high pressure, a high flow rate
decreased the rate of fouling, improving average permeate
flux, but increasing energy consumption. No single optimum
operating condition was found for the set of variables studied.
Membrane Fouling and Concentration Polarization
The main problem in the practical application of membrane
filtration is the reduction in permeate flux over time, caused
by the accumulation of feed components in the membranes porous structure and on the membrane surface, the
chemical interaction between solutes and membrane
material on the membrane surface (concentration polarization and gel formation), and bacterial growth on the
membranes [201, 217]. Concentration polarization is considered reversible and can be controlled in a membrane
module by means of velocity adjustment, pulsation, ultrasound or an electric field. Membrane fouling, on the other
hand, is more complicated in that it is considered as a
group of physical, chemical and biological effects leading
to irreversible loss of membrane permeability [44, 180].
Microbiological fouling of reverse osmosis membranes
is one of the main factors in flux decline and loss of salt
rejection [88]. In investigating the factors affecting NF
fouling materials, several studies have demonstrated that
dissolved organic matters (DOM), mainly humic and fulvic
acid, play a key role in NF [123] and RO flux decline [110,
158]. In order to control membrane fouling during juice
clarification, several flux enhancement methods have been
proposed [207]. The optimal conditions found in each of
the studies were dependent on the juice processed, the
equipment and the procedure. An important parameter
when estimating fouling is to determine, firstly, the
deionized water flux (Jo) that corresponds to the flux of an

123

Food Eng Rev (2011) 3:136158

unfouled membrane. This is subsequently compared with


the flux obtained with the test solution in order to quantify
the reduction in flux and, therefore, the extent of the
fouling. Jo is defined as [184] (Eq. 5):
Jo

gT
Q
g20 C A  DP

where gT is the viscosity of water at temperature T and g20 C


the viscosity of water at 20 C, Q is the clean water flow rate
at temperature T, A is the membrane surface area and DP the
transmembrane pressure difference. This equation is
expressed in (L/m2h) [24] and is valid for dilute solutions.
The relationship between viscosity and temperature is
described by Roorda and van der Graaf [178] and is feed
dependent, although dilute feeds can be described by water.
Flux reduction (FR) with regard to clean water flux can
be determined as a percentage of JOb by comparing the JOa
before and after membrane operation as described by
Manttari and Nystrom [137]:
FRCWF

Job  Joa
 100%
Job

where the indices a and b reflect before and after the filtration of feed, respectively.
Alternatively, flux reduction can also be described as the
difference between permeate and clean water fluxes as
follows:
FRPF

Job  J
 100%
Job

where PF is subscript for permeate flux. To use this


equation, a set pressure for clean water flux and permeate
flux has to be selected and filtration should have reached a
steady state.
The permeate flow rate is directly proportional to the
pressure applied and inversely proportional to the viscosity.
Viscosity can be controlled by two factors: solids concentration in the feed and temperature [176, 189].
Various models have been developed to describe the cake
formation in filtration processes. The resistance-in-series
model (RIS) is derived from the analysis of one-dimensional
Darcian flow through a filter cake. The RIS sums up all
factors influencing the decline in permeate flux, such as
resistance terms, membrane, adsorption, concentration
polarization and fouling resistances [92]. Darcys law is the
basic model used to determine filtration resistance during
permeate transport through porous membranes (Eq. 8) [93]:
JD

DP
gRm

where JD is the permeate flux (L/hm2) the TMP (Pa), DP is


the pressure drop, g the viscosity of the fluids (Pas), and Rm
is the total filtration resistance (m-1) of the membrane.

Food Eng Rev (2011) 3:136158

149

Fig. 6 Schematic of resistances


in series model (adapted from
Chang I et al. [42])

The total resistance, RT, is taken as the sum of each


resistance (RT ? Rm ? Rc ? Rf). This summation is only
possible if each resistance is additive [42]. For an ideal
case as illustrated in Fig. 6, individual resistances act
independently of one another, so that three different
resistances can be added. Cake resistance is certainly a
series resistance, while Rf is defined as the difference
between the fouled membrane resistance (Rm ? Rf) and the
clean membrane resistance, Rm.
A number of semi-empirical and empirical models have
been proposed to describe flux behaviour during membrane
filtration processes. These models can represent flux as a
function of either time or volume concentration ratio [71].
According to these models, change of flux as a function of
volume concentration ratio or time is examined in two or
three periods, characterized by an initial rapid flux decline
followed by a less severe decrease in flux until a steady
state is reached. Exponential models often fit reasonably
well with experimental flux data and have been applied to
membrane filtration of apple juice [149, 192]. Figure 7
shows a typical fouling study protocol in which a new
membrane is first compacted and then the deionized water

Fig. 7 Typical protocol in fouling studies (adapted from Kilduft et al.


[108])

flux is measured. Subsequently, the test solution (in the


case of this figure, a natural organic matter (NOM) solution) is used and the flux is measured [108]. Depending on
the feed water, a significant decline in flux can occur. This
decrease in flux can occur for a number of reasons: (1)
concentration polarization [73], (2) fouling that can be
reversed chemically and (3) irreversible fouling. Concentration polarization can be reversed by water flush,
reversible fouling can be removed with an appropriate
chemical cleaning protocol and irreversible fouling, which
can be due to the irreversible binding of foulants to the
membrane or membrane compaction, cannot be reversed
and will ultimately determine the lifetime of a membrane.
It should be noted that some researchers classify all fouling
that cannot be reversed with a water flush as irreversible
[184].
For constant flux operation, a protocol to determine
fouling would be to measure the variation of transmembrane pressure, which increases with this phenomenon.
Choi et al. [46] subdivided fouling resistance into either
reversible or irreversible fouling and, experimentally,
determined the values of these resistances in the cross-flow
UF of biological suspensions. In UF, the concentration of
solute in the liquid phase and on the cake varies perpendicularly to the membrane or the cake. In recent studies by
Mirsaeedghazi et al. [147], the resistance to permeation due
to this distribution was defined in the concentration
polarization model [86]. Ho et al. [87] demonstrated that
membrane pore morphology can have a significant effect
on the rate of flux decline during protein microfiltration
with membranes having straight-through (non-interconnected) pores, showing a more rapid flux decline than in
membranes with an interconnected pore structure, since the
filtrate can flow under and around any surface blockage
through the interconnected pores [41, 43]. The relationships between permeate flux and resistance are expressed
with equations of the same form, but with different exponents on the permeate flux term for the different types of
blockage mechanisms [104].
Choi et al. [47] proposed a novel fouling index known as
the combined fouling index (CFI), which uses a set of

123

150

different membrane filters to determine the fouling potential of water. The authors suggested that no single method
can be successfully used for accurate prediction of fouling
potential of feed waters, but the combination of different
fouling indices using different types of test membranes
such as hydrophilic MF and hydrophilic UF is a good
possibility, as each of these membranes can capture different portions of foulants in a given feed.
Two types of fouling phenomena are generally distinguished for MF and UF. The first is macrosolute or particle
adsorption, which refers to the specific intermolecular
interactions between the particles and the membrane that
occur even in the absence of filtration. It is usually irreversible, adhesive fouling. In water treatment applications,
the foulants are usually adhesive due to hydrophobic
interactions, hydrogen bonding, van der Waals attractions
and extracellular macromolecular interactions, among
others [123, 131]. The second type is known as filtrationinduced macrosolute or particle deposition, which is often
reversible, non-adhesive fouling, where the accumulation
of cells, pectin and other rejected particles on the top
surface of the membrane is prominent. It occurs as external
fouling or cake formation. Reversible fouling resulting
from cake formation was found to be only weakly dependent on membrane surface chemistry; in contrast, irreversible fouling exhibits a marked dependence on surface
chemistry [197].
Flux Decline
The major limiting factor in membrane filtration processes
is flux decline over time due to membrane fouling,
reducing process efficiency [74, 219]. According to Koros
et al. [112], fouling is a process that results in a membranes decreased effectiveness due to deposition of suspended or dissolved substances on its external surfaces, at
its pore openings, or within its pores. Fouling is also
described as flux decline which is irreversible and can only
be removed by, for example, chemical cleaning [205].
Those flux declines can be reversed with clean water and
are hence not considered as fouling [184]. To minimize
these problems, enzymatic hydrolysis is recommended by
many authors as a pre-treatment for membrane processes in
fruit clarification and concentration [54, 207]. Various
models have been proposed as a means to analyse and
predict flux decline behaviour during the filtration of
macromolecular solutions [173]. All of them can be classified into three broad categories: (a) osmotic pressure
controlled, (b) cake or gel layer controlled and (c) resistance in series models. According to the resistance in series
model, flux decline is due to the combined effects of
irreversible membrane fouling and reversible fouling
(concentration polarization) on the membrane surface in

123

Food Eng Rev (2011) 3:136158

addition to membrane resistance [36, 126]. In the gel


polarization model, the major resistance offered to solvent
flux is resistance of the gel layer over the membrane surface [105, 108]. Accordingly, various attempts to develop
proper pre-treatment processes have been made, including
coagulation, ozonation, chlorination, dissolved air floatation [206] and MF/UF prefiltration. On the other hand, MF
and UF have proved to be attractive methods for the pretreatment of NF feed water [155]. In order to reduce or
eliminate fouling, it is necessary to identify the foulants.
This can be achieved by characterizing the fouled membrane or by performing fouling studies in the laboratory.
Once the foulants are identified, suitable control strategies
can be adapted. An overview of foulants and appropriate
control strategies are summarized in Table 5. The strategies encompass a number of categories [63].
Critical Flux
The critical flux hypothesis is based on the existence of a
flux start-up value below which a decline in flux (or an
increase in TMP) does not occur over time and above
which fouling is observed. According to its definition,
critical flux is influenced by hydrodynamics (pH, ionic
strength, membrane characteristics, particle characteristics,
etc.) and the position on the membrane [76]. Critical flux is
defined as the permeate flux above which an irreversible
deposit (fouling) appears. Below the critical flux, no fouling is observed. Once critical flux is exceeded, the particles
form a stagnant, consolidated and aggregated structure that
is slow to break-up and re-disperse, a process also known
as depolarization [137]. To date, much research has been
undertaken with the aim of assessing the critical flux of a
specific membrane, but most of this work has focused on
ideal situations, in which latex or silica particle suspensions
are employed [117]. Unfortunately, such results cannot be
directly transferred to full-scale applications in real
wastewater treatment due to the complexity of these
matrices. Only a few of experiments have been performed
on submerged membranes with real feed, but in these small
modules have been used with some operational parameters
(i.e. the specific air flow rate for membrane cleaning) far
removed from those typical in full-scale applications [9].
Goosen et al. [73] indicated that in most of the theories that
have been developed, the limiting or critical flux is based
on semi-empirical knowledge rather than being predicted
from fundamental principles. Zhong et al. [222] showed
that there is a critical pressure for a given suspension.
When applied pressure is below that critical pressure, there
will only be a concentration polarization layer on the
membrane surface. A fouling layer, however, will form
between the polarization and the membrane surface when
the applied pressure exceeds the critical pressure. The

Food Eng Rev (2011) 3:136158

151

Table 5 Foulants and their control strategies in nanofiltration and reverse osmosis processes (adapted from Schafer et al. [185])
Foulant

Fouling control

General

Hydrodynamics/shear, operation below critical flux chemical cleaning.

Inorganic
(scaling)

Operate below solubility limit, pre-treatment, reduce pH to 46 (acid addition), low recovery, additives (antiscalants).

Organics

Pre-treatment using biological processes (enzymatic treatment), activated carbon, ion exchange (e.g. MIEX), ozone,
enhanced coagulation.
Pre-treatment using coagulation and filtration, microfiltration, ultrafiltration.

Colloids
(\0.5 lm)
Biological solids

Enzymatic treatment using ultrafiltration, microfiltration.


Pre-treatment using disinfection (e.g. chlorination/dechlorination), filtration, coagulation, microfiltration.

limiting or critical flux values predicted by the mechanistic


model should therefore be applied to operate systems at
just below the determined critical flux. Cross-flow velocity
increases critical flux while solute concentration decreases
it. Repulsion between solute and membrane also increases
critical flux in the case of high molar mass polysaccharides,
while in paper industry effluents only weak forms of critical flux were found [184]. Operators of RO/UF plants/
units should therefore operate their systems just below the
critical flux in order to maximize productivity while minimizing membrane fouling.
Membrane Cleaning
The cleaning process, including cleaning frequency,
backwash time, etc., must therefore be optimized in order
to minimize the adverse effects of cleaning agents on
membrane life, the cost of purchasing and disposing of
cleaning chemicals, the volumes of water consumed and
wastewater produced, and the disruption to normal manufacturing operations [51, 132]. However, investment in
purchasing and replacing membranes in industrial plants is
still a very significant cost. Therefore, membrane cleaning
is an essential step in maintaining the performance of the
membrane process, and as such, it has been the subject of a
large amount of research work [4]. The cleaning process
must remove fouling deposits and restore the normal
capacity (measured, usually, as demineralized water flux)
and separation characteristics of the equipment [136]. In
practice, this is achieved through a variety of chemical and
physical interactions between the cleaning solution and the
soil on the membrane surface [23]. Reverse osmosis permeate may also be of suitable quality for use in cleaning
[73].
Membranes are currently cleaned using a combination
of water and air in either the forward or backward direction. Effective cleaning after juice processing involves
rinsing the membrane system with water initially and after
each cleaning step. Membrane manufacturers generally
recommend the use of high-quality water such as filtered

and demineralized water [13]. Installing and running water


purification systems, however, is expensive. Alternatively,
water treatment with chemicals such as sequestering agents
(e.g. EDTA, polyphosphates) can be added to low-quality
water to increase the solubility of metal ions such as calcium, magnesium, manganese and iron [202]. When the
abovementioned cleaning methods are not effective enough
to restore the flux to an acceptable level, it is necessary to
clean the membranes chemically. During chemical cleaning,
membranes are soaked in a solution of strong acids and bases
such as hydrogen chloride (HC1) or sodium hydroxide
(NaOH), or disinfecting agents such as hypochlorous acid
(HOC1). Effective chemical cleaning restores the initial
flux, making the membrane serviceable for further operation
[114, 115].
The cleaning agents used have different functions, such
as the emulsification and degradation of fat and proteins,
disinfection, removal of fat and inorganic particles. [19].
Seng Ang et al. [10] found that a metal chelating agent
(EDTA) and an anionic surfactant (SDS) were able to clean
the fouled RO membrane effectively by optimizing
chemical (dose and pH) and physical (time, cross-flow
velocity, and temperature) conditions during cleaning. The
permeate flux was, however, poorly recovered when
cleaning was performed with NaOH (pH 11). Interfacial
force measurements (by atomic force microscopy) further
confirmed the cleaning results, demonstrating the influence
of cleaning agents on reducing the foulantfoulant adhesion force. The results showed that the adhesion force
caused by calcium-induced foulantfoulant interaction was
significantly reduced with EDTA and SDS cleaning, while
substantial adhesion force still remained following NaOH
cleaning.
Cleaning protocols may be adequate to remove membrane foulants, but restoring permeate flux is not necessarily an indication that the system is completely sanitized
[159]. Microorganisms may remain on various surfaces
(e.g. membrane, housing, antitelescopic devices, etc.) after
cleaning/sanitization cycles. An important research effort
has been made in recent years to design and optimize

123

152

cleaning protocols in order to regenerate ceramic membranes after the filtration of a wide number of materials
[219]. There are many different cleaning regimes, and the
exact procedure for a given membrane system depends on
the product treated, the membrane type and the system
design [209].
Three cleaning methods can be distinguishedhydraulic, mechanical and chemical cleaningand the choice
depends on the module configuration, the chemical resistance of the membrane and the type of foulant encountered.
Hydraulic cleaning methods include backflushing,
which consists of alternating pressurizing and depressurizing by changing the flow direction at a given frequency.
Backflushing is an in situ method of cleaning the membrane by periodically reversing the transmembrane flow. In
this way, the stationary concentration polarisation profiles
are disturbed and the fouling layer is removed from inside
the membrane and from the membrane surface. The
backflushing medium can be the permeate, another liquid
or gas, but if the permeate flow is used for flushing, it
results in a loss of permeate against an increased flux [194].
Mechanical cleaning can only be applied in tubular
systems using oversized sponge balls [51]. Physical
cleaning generally uses mechanical forces to remove foulants from the membrane surface [124] and such methods
include backflush/forward flush/reverse flush, scrubbing
(e.g. using foam balls for tubular modules), air sparging,
CO2 back permeation, vibration and sonication. Sonication
is a relatively novel method for membrane cleaning,
although ultrasound is commonly used in membrane
autopsies to remove the fouling deposits from the membranes for chemical analysis [184]. Lamminen et al. [118]
indicated that ceramic membranes may be effectively
cleaned using ultrasound at frequencies from 70 to
620 kHz without damage to the membranes. Increases in
the power intensity of the system were found to increase
the cleaned flux ratio. This increase was attributed to an
increase in the number of cavitation bubbles in the system
and an increase in acoustic energy in the system.
Chemical cleaning is the most common method of
reducing fouling, with a number of chemicals being used
separately or in combination. The concentration of the
chemicals and the compounds used are key parameters that
should be chosen in keeping with the chemical resistance
of the membrane. The frequency with which membranes
need to be cleaned can be estimated from a process optimization step [134].
The cleaning procedure employed by Martinez-Ferez
et al. [139] to regenerate the membranes after operation
was as follows: (1) rinsing with demineralized water, (2)
recirculating a solution of 20 g/L sodium hydroxide and
0.1 g/L sodium dodecyl sulphate at 50 C for 30 min and
(3) rinsing with demineralized water until neutrality was

123

Food Eng Rev (2011) 3:136158

achieved. Membranes were considered to be regenerated


after the resistance of the clean membrane was recovered.

Final Remarks
The potential advantages of membrane filtration techniques
over conventional filtration techniques in fruit juice processing are undeniable and include improved product
quality, easy scaling up of production and lower energy
consumption. However, these techniques are generally
limited by problems related to fouling and by the relatively
short lifespan of the membranes. In general, well-developed process engineering (including pre-treatment) is of
fundamental importance for the success of the membrane
process. Advanced membrane and module materials need
to be matched with the appropriate economic manufacturing processes.
The use of membrane processes associated with enzymatic hydrolysis results in clarified and concentrated fruit
juices with high nutritional and sensory quality. Clarification considerably reduces the amount of suspended solids
and the turbidity of fresh juice. An additional advantage of
clarification is that it allows juice to be processed in a
single step, reducing working times and increasing yield in
terms of the volumes of clarified juice produced. The
concentration of fruit juices by reverse osmosis facilitates
handling of the product during transport and storage
without negatively affecting its sensory and nutritional
properties.
Industrial applications have recently been developed for
fruit, vegetable and sugar juices. Continued efforts to
develop improved membrane materials, modules and process designs should enable membrane systems to play an
important role in the next generation of biotechnology
processes. The further integration of membrane operations
is expected, provided they are designed in such a way that
at each processing step, end products, by-products and
waste products are given equal attention. The safety and
quality of the products manufactured must be ensured with
regard to microbiological, functionality, texture, flavour
and taste factors.
Future trends for the use of membranes in the fruit juice
industry will be driven by the quest to achieve higher
selectivity and permeability in clarification processes, to
boost process intensification and to decrease operating
costs as well as the costs of membrane production. There is
a trend towards increasing the use of disposable systems
(bioreactors, ultrafilter membranes and enzymatic membranes) in process intensification which are attractive for
production scale manufacturing, eliminating the need
for the development and validation of cleaning cycles.
Future developments will determine whether such

Food Eng Rev (2011) 3:136158

membrane-based processes can provide the required


product quality, purity, yield and throughput while
remaining economically viable for the fruit juice industry.
Acknowledgments A.P. Echavarra thanks the Commission for
Universities and Research of the DIUE of the Generalitat de
Catalunya and the European Social Fund for the predoctoral grant.

References
1. Abbas H, Hossain MM, Chen XD (2006) A laboratory investigation of the anhydrous milk fat fractionation using a membrane
technique. Sep Purif Tech 48:167175
2. Adams A, Borrelli C, Fogliano V, De Kimpe N (2005) Thermal
degradation studies of food melanoidins. J Agric Food Chem
53(10):41364142
3. Alkorta I, Garbisu C, Llama M, Serra J (1995) Viscosity
decrease of pectin and fruit juices catalyzed by pectin lyase from
Penicillium italicum in batch and continuous-flow membrane.
Biotechnol Tech 9(2):95100
4. Almecija M, Zapata JE, Martinez-Ferez A, Guadixa A,
Hernandez A, Calvo J, Guadixa EM (2009) Analysis of cleaning
protocols in ceramic membranes by liquidliquid displacement
porosimetry. Desalination 246:168172
5. Alper N, Bahceci SK, Acar J (2005) Influence of processing and
pasteurization on color values and total phenolic compounds of
pomegranate juice. J Food Process Pres 29:357368
6. Alvarez S, Riera F, Alvarez R, Coca J, Cuperus FP, Bouwer S,
Boswinkel G, van Gemert RW, Veldsink JW, Giorno L, Donato
L, Todisco S, Drioli E, Olsson J, Tragardh G, Gaeta SN, Panyor
L (2000) A new integrated membrane process for producing
clarified apple juice and apple juice aroma concentrate. J Food
Process Eng 46(2):109125
7. Alvarez S, Riera FA, Alvarez R, Coca J (2002) Concentration of
apple juice by reverse osmosis at laboratory and pilot-plant
scales. Ind Eng Chem Res 41:61566164
8. Amuda O, Amoob IA, Ajayi O (2006) Performance optimization
of coagulant/flocculant in the treatment of wastewater from a
beverage industry. J Hazard Mater 129:6972
9. Andreottola G, Guglielmi G (2003) Critical flux determination
in two MBRs for municipal wastewater treatment. In: Proc of
IMSTEC03 1014, Sydney
10. Ang SW, Lee S, Elimelech M (2006) Chemical and physical
aspects of cleaning of organic-fouled reverse osmosis membranes. J Memb Sci 272:198210
11. Aporn L, Zhenyu L, Sasitorn T, Suphitchaya C, Wirote Y (2010)
Effect of membrane property and operating conditions on phytochemical properties and permeate flux during clarification of
pineapple juice. J Food Eng 100:514521
12. Araujo WA, Maciel MR (2005) Reverse osmosis concentration
of orange juice using spiral wound membranes. Alim Nutr
Araraquara 16(3):213219
13. Astudillo C, Parra J, Gonzalez S, Cancino B (2010) A new
parameter for membrane cleaning evaluation. Sep Purif Technol
73(2):286293
14. Bagger-Jrgensen R, Meyer A (2004) Effects of different
enzymatic pre-press maceration treatments on the release of
phenols into blackcurrant juice. Chem Mater Sci 219(6):620
629
15. Bagger-Jrgensen R, Meyer A, Pinelo M, Varming C, Jonsson G
(2011) Recovery of volatile fruit juice aroma compounds by
membrane technology: sweeping gas versus vacuum membrane
distillation. IFSET. doi:10.1016/j.ifset.2011.02.005

153
16. Barrett DM, Somogyi L, Ramaswamy H (2005) Processing
fruits. Science and technology, 2nd edn. CRC Press, Florida
17. Barros S, Mendes E, Peres L (2004) Influence of depectinization
in the ultrafiltration of West Indian cherry (Malpighia glabra L.)
and pineapple (Ananas comosus juices). Cienc Tecnol Aliment
24(2):194201
18. Behnaz R, Abdolreza A, Ahmadreza R, Mahdi F (2011) Clarification of tomato juice by cross-flow microfiltration. Int J Food
Sci Tech 46:138145
19. Bejar J, Cortes U (2001) Agent-mediated electronic commerce
III agent strategies on DPB auction tournaments. Springer,
London
20. Benitez E, Lozano JE (2007) Effect of gelatin on apple juice
turbidity. Lat Am Appl Res 37:261266
21. Biscaro D, Cano E (2010) Purification and characterization of
the exopoly galacturonase produced by Aspergillus giganteus in
submerged cultures. J Ind Microbiol Biotechnol 37:567573
22. Bitange NT, Zhang W, Shi-Ying X, Wenbin Z (2009) The
Influence of a pectinase and pectinase/hemicellulases enzyme
preparations on percentage pineapple juice recovery, particulates and sensory attributes. Pak J Nutr 8(8):11841189
23. Blanpain-Avet P, Migdal F, Benezech T (2009) Chemical
cleaning of a tubular ceramic microfiltration membrane fouled
with a whey protein concentrate suspensioncharacterization of
hydraulic and chemical cleanliness. J Memb Sci 337:153174
24. Boerlage S, Kennedy MD, Dickson M, El-Hodali EY, Schippers
JC (2002) The modified fouling index using ultrafiltration
membranes (MFI-UF): characterisation, filtration mechanisms
and proposed reference membrane. J Memb Sci 197:121
25. Bonnelye V, Guey L, Del Castillo J (2008) UF/MF as RO pretreatment: the real benefit. Desalination 222:5965
26. Borneman Z, Gtkmen V, Herry Nijhuis H (1997) Selective
removal of polyphenols and brown colour in apple juices using
PES/PVP membranes in a single-ultrafiltration process. J Memb
Sci 134:191197
27. Bouchoux A, Roux-de Balmann H, Lutin F (2006) Investigation
of nanofiltration as a purification step for lactic acid production
processes based on conventional and bipolar electrodialysis
operations. Sep Purif Tech 52(2):266273
28. Bowen WR, Doneva TA, Yin HB (2002) Atomic force
microscopy studies of membranesolute interactions (fouling).
Desalination 146:97102
29. Brummell D, Harpster M (2001) Cell wall metabolism in fruit
softening and quality and its manipulation in transgenic plants.
Plant Mol Biol Rep 47:311340
30. Burdulu HS, Karadeniz F (2003) Effect of storage on nonenzymatic browning of apple juice concentrates. Food Chem
80(1):9197
31. Carrin ME, Ceci LN, Lozano JE (2004) Characterization of
starch in apple juice and its degradation by amylases. Food
Chem 87:173178
32. Carrn M, Ceci L, Lozano JE (2001) Effects of co-immobilization of pectinase and amylase on ultrafiltration of apple juice
simulate. JFPE 24(6):423435
33. Carvalho LM, Bento da Silva CA (2010) Clarification of pineapple juice by microfiltration. Cienc Tecnol Aliment 30(3):828
832
34. Carvalho L, Castro I, Da Silva BC (2008) A study of retention of
sugars in the process of clarification of pineapple juice (Ananas
comosus, L. Merril) by micro and ultra-filtration. J Food Eng
87:447454
35. Cassano A, Donato L, Drioli E (2007) Ultrafiltration of kiwifruit
juice: operating parameters, juice quality and membrane fouling.
J Food Process Eng 79:613621
36. Cassano A, Marchio M, Drioli E (2007) Clarification of blood
orange juice by microfiltration: analyses of operating

123

154

37.

38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51.
52.

53.

54.

55.

Food Eng Rev (2011) 3:136158


parameters, membrane fouling and juice quality. Desalination
212(13):1527
Cassano A, Tasselli F, Conidi C, Drioli E (2009) Ultrafiltration
of clementine mandarin juice by hollow fibre membranes.
Desalination 241:302308
Cassano A, Conidi C, Drioli E (2010) Physico-chemical
parameters of cactus pear (Opuntia ficus-indica) juice clarified
by microfiltration and ultrafiltration processes. Desalination
250:11011104
Ceci L, Lozano J (1998) Determination of enzymatic activities
of commercial pectinases for the clarification of apple juice.
Food Chem 61(12):237241
Chaisakdanugull C, Theerakulkait C, Wrolstad R (2007) Pineapple juice and its fractions in enzymatic browning inhibition of
banana [Musa (AAA Group) Gros Michel]. J Agric Food Chem
55:42524257
Chandler M, Zydney A (2006) Effects of membrane pore
geometry on fouling behavior during yeast cell microfiltration.
J Memb Sci 285(12):334342
Chang IS, Field R, Cui Z (2009) Limitations of resistance-inseries model for fouling analysis in membrane bioreactors: a
cautionary note. DWT 8:3136
Chia-Chi H, Zydney A (2006) Overview of fouling phenomena
and modeling approaches for membrane bioreactors. Sep Sci
Tech 41(7):12311251
Chien-Hwa Y, Lung-Chen F, Shaik Khaja L, Chung-Hsin W,
Cheng-Fang L (2010) Enzymatic treatment for controlling
irreversible membrane fouling in cross-flow humic acid-fed
ultrafiltration. J Hazard Mater 177:11531158
Chisari M, Barbagallo RN, Spagna G (2007) Characterization of
polyphenol oxidase and peroxidase and influence on browning
of cold stored strawberry fruit. J Agric Food Chem 55(9):
34693476
Choi H, Zhang K, Dionyosiou D, Oerther D, Sorial G (2005)
Influence of cross-flow velocity on membrane performance
during filtration of biological suspension. J Memb Sci 248:189
191
Choi JS, Hwang TM, Lee S, Hong S (2009) A systematic
approach to determine the fouling index for a RO/NF membrane
process. Desalination 238(13):117127
Chornomaz PM, Ochoa NA, Pagliero C, Marchese J (2011)
Synthesis, characterization and performance of membranes for
clarification of lemon juice. DWT 27(13):294298
Cleveland C, Seacord T, Zander A (2002) Standardized membrane pore size characterization by polyethylene glycol rejection. J Environ Eng 128(5):399
Curreli N, Rinaldi A (1998) Effect of 3-hydroxyanthranilic acid
on mushroom tyrosinase activity. Protein Struct Mol Enzymol
1384(2):268276
Dsouza N, Mawson AJ (2005) Membrane cleaning in the dairy
industry: a review. Crit Rev Food Sci Nutr 45(2):125134
Damasceno L, Fernandes F, Magalhaes M, Brito E (2008)
Evaluation and optimization of non enzymatic browning of cajuina during thermal treatment. Braz J Chem Eng 25(2):313320
Daufin G, Escudier J, Carre`re H, Bero`t S, Fillaudeau L, Decloux
M (2001) Recent and emerging applications of membrane processes in the food and dairy industry. FBP 79(2):89102
De Bruijn JA, Venegas JA, Martinez R, Borquez R (2003)
Ultrafiltration performance of Carbosep membranes for the
clarification of apple juice. LWT 36:397406
De Roos A, Grassin C, Herweijer M, Kragh K, Poulsen C, Soe J,
Sorensen J, Wilms J (2004) Industrial enzymes: enzymes in food
applications. In: W. Aehle (ed) Enzymes in industry: production
and applications, 2nd edn. Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim. doi:10.1002/3527602135.ch5a

123

56. Diano N, Grimaldi T, Bianco M, Rossi S, Gabrovska K,


Yordanova G, Godjevargova T, Grano V, Nicolucci C, Mita L,
Bencivenga U, Canciglia P (2008) Apple juice clarification by
immobilized pectolytic enzymes in packed or fluidized bed
reactors. J Agric Food Chem 56:1147111477
57. Diban N, Voinea OC, Urtiaga A, Ortiz I (2009) Vacuum
membrane distillation of the main pear aroma compound:
experimental study and mass transfer modeling. J Memb Sci
326:6475
58. Diqiao S, Wei M, Hossain Z, Saleh S (2007) Separation of
polyphenolics and sugar by ultrafiltration: effects of operating
conditions on fouling and diafiltration. WASET 34:1724
59. Drioli E, Curcio E (2007) Membrane engineering for process
intensification: a perspective. J Chem Technol Biotechnol 82:
223227
60. Dyson P, Ranking R (2008) Fluid properties at nano/meso scale:
a numerical treatment. John Wiley & Sons, Ltd., Chichester,
pp 182184
61. Echavarria A, Pagan J, Ibarz A (2011) Effect of previous
enzymatic recirculation treatment through a tubular ceramic
membrane on ultrafiltration of model solution and apple juice.
J Food Process Eng 102(4):334339
62. Eissa H, Fadelb H, Ibrahim GE, Hassan I, Elrashid AA (2006)
Thiol containing compounds as controlling agents of enzymatic
browning in some apple products. Food Res Int 39(8):855863
63. Fane AG, Beatson P, Li H (2000) Membrane fouling and its
control in environmental applications. Water Sci Technol
41(1011):303308
64. Feng-Lin G, Moon J, Abbas S, Xiao-Ming Z, Shu-Qin X,
Zheng-Xing C (2010) Structure and antioxidant activity of high
molecular weight Maillard reaction products from caseinglucosa. Food Chem 120:505511
beda F, Cordero-Otero C, Thanvan65. Fernandez-Gonzalez M, U
thri Gururajanb V, Brionesa A (2005) Engineering of an oenological Saccharomyces cerevisiae strain with pectinolytic
activity and its effect on wine. Int J Food Microbiol 102(2):173
183
66. Galaverna G, Di Silvestro G, Cassano A, Sforza S, Dossena A,
Drioli E, Marchelli R (2008) A new integrated membrane process for the production of concentrated blood orange juice:
effect on bioactive compounds and antioxidant activity. Food
Chem 106:10211030
67. Galeazzi M, Sgarbieri V, Conctantinides S (1981) Isolation,
purification and physicochemical characterization of polyphenol
oxidases (PPO) from a dwarf variety of banana (Musa cavendishii L.). J Food Sci 46:150155
68. Garca-Martn N, Perez-Magarin S, Ortega-Heras M, GonzalezHuerta C, Mihnea M, Gonzalez-Sanjose ML, Palacio L,
Pradanos P, Hernandez A (2010) Sugar reduction in musts with
nanofiltration membranes to obtain low alcohol-content wines.
Sep Purif Tech 76:158170
69. Garza S, Ibarz A, Pagan J (1999) Non-enzymatic browning in
peach puree during heating. Food Res Int 32:335343
70. Girard B, Fukumoto LR (2000) Membrane processing of fruit
juices and beverages: a review. Crit Rev Biotechnol 20(2):109
175
(2007) Effect of pretreatment with
71. Gokmen V, Cetinkaya O
gelatin and bentonite on permeate flux and fouling layer resistance during apple juice ultrafiltration. J Food Eng 80:300305
72. Gokoglu N, Yerlikaya P (2008) Inhibition effects of grape seed
extracts on melanosis formation in shrimp (Parapenaeus
longirostris). IJFST 43:10041008
73. Goosen MF, Sablani SS, Roque-Malherbe R (2005) Membrane
fouling: recent strategies and methodologies for its minimization. In: Handbook of membrane separations: chemical,

Food Eng Rev (2011) 3:136158

74.

75.
76.

77.

78.

79.

80.

81.
82.

83.

84.

85.

86.

87.

88.

89.

90.

91.

92.

93.
94.

pharmaceutical, and biotechnological applications. Marcel


Dekker, New York (in press)
Guadix A, Zapata JE, Almecija MC, Guadix EM (2010) Predicting the flux decline in milk cross-flow ceramic ultrafiltration
by artificial neural networks. Desalination 250:11181120
Guang-Yuan Z, Wei Z, Guang-Jie A (2008) Effect of storage on
cloud stability of cloudy apple juice. Food Sci Technol Int 14:105
Guglielmia G, Chiarania D, Juddb S, Andreottola G (2007) Flux
criticality and sustainability in a hollow fibre submerged membrane bioreactor for municipal wastewater treatment. J Memb
Sci 289(12):241248
Gurak PD, Cabral LM, Rocha-Leao MH, Matta VM, Freitas SP
(2010) Quality evaluation of grape juice concentrated by reverse
osmosis. J Food Process Eng 96:421426
Gyura J, Seres Z, Vatai G, Bekassy Molnar E (2002) Separation
of non-sucrose compounds from the syrup of sugar-beet processing by ultra- and nanofiltration using polymer membranes.
Desalination 148:4956
Haitao L, Ka-Wing C, Chi-Hin C, Zhendan He, Mingfu W (2007)
Oxyresveratrol as an antibrowning agent for cloudy apple juices
and fresh-cut apples. J Agric Food Chem 55(7):26042610
Hakimzadeh V, Razavi S, Piroozifard M, Shahidi M (2006) The
potential of microfiltration and ultrafiltration process in purification of raw sugar beet juice. Desalination 200:520522
Hande S, Feryal K (2003) Effect of storage on nonenzymatic
browning of apple juice concentrates. Food Chem 80:9197
He Q, Luo Y, Chen P (2008) Elucidation of the mechanism of
enzymatic browning inhibition by sodium chlorite. Food Chem
110:847851
Hea Y, Zhijuan J, Lia Shunxin (2007) Effective clarification of
apple juice using membrane filtration without enzyme and
pasteurization pretreatment. Sep Purif Tech 57(2):366373
Hidalgo F, Nogales F, Zamora R (2003) Effect of the pyrrole
polymerization mechanism on the antioxidative activity of
nonenzymatic browning reactions. J Agric Food Chem 51:
57035708
Hitoshi M, Susumu A, Koji S, Yasunari K (2000) Water recycling by floating media filtration and nanofiltration at a soft
drink factory. Desalination 13(1):4753
Ho C, Zydney A (2001) Protein fouling of asymmetric and
composite microfiltration membranes. Ind Eng Chem Res
40:14121421
Ho C, Zydney A (2003) Effect of membrane morphology on
system capacity during normal flow microfiltration. Biotech
Bioeng 83(5):537543
Hoeka E, Allredb J, Knoellc T, Byeong-Heon J (2008) Modeling
the effects of fouling on full-scale reverse osmosis processes.
J Memb Sci 314(12):3349
Holanda CM, Herculano PN, Souza Porto T, Simas T, Aparecida
M, De Souza-Motta CM (2011) Production and partial characterization of pectinases from forage palm by Aspergillus niger
URM4645. A J B 10(13):24692475
Hoondal G, Tiwari R, Tewari R, Dahiya N, Beg Q (2002)
Microbial alkaline pectinases and their industrial applications: a
review. Appl Microbiol Biotechnol 59:409418
Hussain AA, Al-Rawajfeh AE (2009) Recent patents of nanofiltration applications in oil processing. Desalination, wastewater
and food industries the rheology of semiliquid foods. Recent Pat
Chem Eng 2:5166
Hyeok C, Zhang Kai, Dionysios D, Oesther D, Serial G (2005)
Effect of permeate flux and tangential flow on membrane fouling for wastewater treatment. Sep Purif Technol 45(1):6878
Ibarz A, Barbosa-Canovas GV (2003) Unit operations in food
engineering. CRC Press, Florida, pp 288293
Ibarz A, Barbosa-Canovas GV (2005) Operaciones unitarias en
la ingeniera de alimentos. Mundi-Prensa, Madrid, pp 318325

155
95. Ibarz I, Gonzales C, Esplugas S, Miguelsanz R (1990) Nonenzymatic browning kinetics of clarified peach juice at different
temperatures. Confructa 34:152159
96. Ibarz A, Pagan J, Garza S (1999) Kinetic models for colour
changes in pear puree during heating at relatively high temperaturas. J Food Eng 39(4):415422
97. Jaffrin M, Lu-Hui D, Akoum O, Brou A (2004) A hydrodynamic
comparison between rotating disk and vibratory dynamic filtration systems. J Food Eng 242(12):155167
98. Jesus DF, Leite MF, Silva LF, Modesta RD, Matta VM, Cabral
LM (2007) Orange (Citrus sinensis) juice concentration by
reverse osmosis. J Food Process Eng 81:287291
99. Jiao B, Cassano A, Drioli E (2004) Recent advances on membrane processes for the concentration of fruit juices: a review.
J Food Process Eng 63:303324
100. Jiaping PC, Honghui M, Lawrence KW, Takeshi M, Yuting W
(2008) Membrane separation: basics and applications. Memb
Desal Tech 13:271332
101. IFFJP (1985) International federation of fruit juice producers
methods. Analysen-analyses, vol 12. Fruit-Union Suisse Assoc,
Suizzera Frutta, pp 12
102. Kajitvichyanukul P, Shammas NK, Yung-Tse H, Lawrence
KW, Jirapat A (2008) Potable water biotechnology, membrane filtration and biofiltration. Memb Desal Tech 13:477
523
103. Kammerer J, Kammerer D, Reinhold C (2010) Impact of saccharides and amino acids on the interaction of apple polyphenols
with ion exchange and adsorbent resins. J Food Eng 98(2):
230239
104. Karasu K, Yoshikawa S, Ookawara S, Ogawa K, Kentish S,
Stevens G (2010) A combined model for the prediction of the
permeation flux during the cross-flow ultrafiltration of a whey
suspension. J Memb Sci 361:7177
105. Karode SKA (2000) Method for prediction of the gel layer
concentration in macromolecular ultrafiltration. J Memb Sci
171:131139
106. Kashyap D, Vohra P, Chopra S, Tewari R (2001) Applications
of pectinases in the commercial sector: a review. Bioresour
Technol 77:215227
107. Katasonova ON, Fedotov PS (2009) Methods for continuous
flow fractionation of microparticles: outlooks and fields of
application. J Anal Chem 64(3):212225
108. Kilduft JE, Mattaraj SB, Pieracci J, Belfort G (2000) Photochemical modification of poly(ether sulfone) and sulfonated
poly(sulfone) nanofiltration membranes for control of fouling by
natural organic matter. Desalination 132:133142
109. Kochubovskiu M (2007) Use of membrane filtration for water
treatment with examples from the republic of Macedonia.
NATO Sec Sci 1(39):193205
110. Kotuniewicz A (2007) Importance of membranes in clean
technologies. Chem Process Eng 29:2941
111. Komthong P, Katoh T, Igura N, Shimoda M (2006) Changes in
the odours of apple juice during enzymatic browning. Food Qual
Prefer 17(6):497504
112. Koros WJ, Ma YH, Shimizu T (1996) Terminology for membranes and membrane processesIUPAC recommendations.
J Memb Sci 120:149159
113. Kozak A, Banvolgyi S, Vincze I, Kiss I, Bekassy-Molnar E,
VataI G (2008) Comparison of integrated large scale and laboratory scale membrane. Chem Eng Process 47:11711177
114. Kuzmenko D, Arkhangelsky E, Belfer S, Freger V, Gitis V
(2004) Chemical cleaning of UF membranes fouled by BSA.
Desalination 179(13):323333
115. Kuzmenko D, Arkhangelsky E, Belfer S, Freger V (2005) Vital
chemical cleaning of UF membranes fouled. Desalination
179:323333

123

156
116. Kwang-Sup Y, Joo-Heon H, Dong-Ho B, Seok-Joong K, SoonDong K (2004) Effective clarifying process of reconstituted
apple juice using membrane filtration with filter-aid pretreatment. J Memb Sci 228:179186
117. Kwon D, Vigneswaran S, Fane A, Ben R (2000) Experimental
determination of critical flux in cross-flow microfiltration. Sep
Purif Tech 19:169181
118. Lamminen M, Walker H, Weavers L (2004) Mechanisms and
factors influencing the ultrasonic cleaning of particle-fouled
ceramic membranes. J Memb Sci 237:213223
119. Landbo A, Meyer A (2004) Effects of different enzymatic
maceration treatments on enhancement of anthocyanins and
other phenolics in black currant juice. IFSET 5:503513
120. Laska ME, Brooks RP, Gayton M, Pujar NS (2005) Robust
scale-up of dead end filtration: impact of filter fouling mechanisms and flow distribution. Biotechnol Bioeng 92(3):308320
121. Lawrence KW, Shammas NK, Cheryan M, Yu-Ming Z, ShuaiWen Z (2008) Treatment of food industry foods and wastes by
membrane filtration. Memb Desal Tech 13:237269
122. Le Clech P, Jefferson B, Chang IS, Judd SJ (2003) Critical flux
determination by the flux-step method in a submerged membrane bioreactor. J Memb Sci 227:8193
123. Lee Sangho (2006) Microfiltration and ultrafiltration as a pretreatment for nanofiltration of surface water. Sep Sci Tech
41:123
124. Lianga H, Gong W, Chen J, Lia G (2008) Cleaning of fouled
ultrafiltration (UF) membrane by algae during reservoir water
treatment. Desalination 220:267272
125. Lisitsin D, Hasson D, Semiat R (2005) Critical flux detection in
a silica scaling RO system. Desalination 186:311318
126. Lopez F, Pescador P, Guell C, Morales M, Garca-Parrilla M,
Troncoso AM (2005) Industrial vinegar clarification by crossflow microfiltration: effect on colour and polyphenol content.
J Food Eng 68(1):133136
127. Lozano JE (2006) Fruit manufacturing: scientific basis, engineering properties, and deteriorative reactions of technological
importance (Food Engineering Series). Springer, USA,
pp 183184, 198199
128. Lozano JE, Drudis-Biscarri R, Ibarz-Ribas A (1994) Enzymatic
browning in apple pulps. J Food Sci 59(3):564567
129. Lukanin OS, Gunko SM, Bryk MT, Nigmatullin R (2003) The
effect of content of apple juice biopolymers on the concentration
by membrane distillation. J Food Eng 60(3):275280
130. Luo Y, Lua S, Zhoua B, Feng H (2011) Dual effectiveness of
sodium chlorite for enzymatic browning inhibition and microbial inactivation on fresh-cut apples. LWT 44(7):16211625
131. Ma HM, Bowman CN, Davis RH (2000) Membrane fouling
reduction by backpulsing and surface modification. J Memb Sci
173:191200
132. Madaeni S, Mansourpanah Y (2004) Chemical cleaning of
reverse osmosis membranes fouled by whey. Desalination
161:1324
133. Madaeni SS, Zereshki S (2010) Energy consumption for sugar
manufacturing. Part I: evaporation versus reverse osmosis.
Energy Convers Manag 51:12701276
134. Madaeni S, Toraj M, Kazeni M (2001) Chemical cleaning of
reverse osmosis membranes. Desalination 134(13):7782
135. Madaeni SS, Naghdi Sedeh S, De Nobili M (2006) Ultrafiltration of humic substances in the presence of protein and metal
ions. Transp Porous Media 65:469484
136. Madaeni S, Rostami E, Rahimpour A (2010) Surfactant cleaning
of ultrafiltration membranes fouled by whey. Int J Dairy Technol
63(2):273283
137. Manttari M, Nystrom M (2000) Critical flux in NF of high molar
mass polysaccharides and effluents from the paper industry.
J Memb Sci 170(2):257273

123

Food Eng Rev (2011) 3:136158


138. Manttari M, Puro L, Nuortila-Joniken J, Nystrom M (2000)
Fouling effects of polysaccharides and humic acid in nanofitration. J Memb Sci 165:117
139. Martinez-Ferez A, Zapata JE, Guadix A, Almecija M, Gomez
CM, Guadix EM (2009) Obtention of goat milk permeates
enriched in lactose-derived oligosaccharides. Desalination 24:
730736
140. Martins S, Jongen W, Van Boekel M (2001) Maillard reaction in
food and implications to kinetic modeling. Trends Food Sci
Tech 11:364373
141. Mascan M (2001) Kinetics of colour change of kiwifruits during
hot air and microwave drying. J Food Eng 48(2):169175
142. Maskan A, Kaya S, Maska M (2002) Effect of concentration and
drying processes on color change of grape juice and leather
(pestil). J Food Eng 54755480
143. Mastromatte M, Conte A, Del Nobile M (2009) Preservation of
fresh-cut produce using natural compounds. Stewart Postharvest
Rev 4(4):17
144. Matta VM, Moretti RH, Cabral LMC (2004) Microfiltration and
reverse osmosis for clarification and concentration of acerola
juice. J Food Process Eng 61:477482
145. Meng X, Slominski BA, Nyachoti CM, Campbell LD, Guenter W
(2005) Degradation of cell wall polysaccharides by combinations
of carbohydrase enzymes and their effect on nutrient utilization
and broiler chicken performance. Poult Sci 84(1):3747
146. Minussi R, Pastore G, Duran N (2002) Potential applications of
laccase in the food industry. Review. Trends Food Sci Tech
13(67):205216
147. Mirsaeedghazi H, Mousavi S, Emam-Djomeh, Rezaei K, Aroujalian A, Navidbakhsh M, Seyyedeh M, Emam-Djomeh Z,
Rezaei K (2011) Comparison between ultrafiltration and microfiltration in the clarification of pomegranate juice. J Food
Process Eng 28. doi:10.1111/j.1745-4530.2010.00598
148. Missang E, Denes M, Baron A (1996) Study of diffusion and
adsorption processes during the maceration of apple tissue by
polygalacturonases. Food Biotechnol 13(1):6792
149. Mondor M, Girard B, Moresoli C (2000) Modeling of flux
behavior for membrane filtration of apple juice. Food Res Int
33:539548
150. Mulder M (1996) Basic principles of membrane technology.
Kluwer Academic Press, Dordrecht, The Netherlands
151. Muratore G, Licciardello F, Restuccia C, Puglisi M, Giudici P
(2006) Role of Different Factors Affecting the Formation of
5-Hydroxymethyl-2-furancarboxaldehyde in Heated Grape
Must. J Agric Food Chem 54:860-863
152. Myung-man Kim, Andrew L. Zydney (2005) Particleparticle
interactions during normal flow filtration: Model simulations.
Chem Engin Scien 60: 4073 4082
153. Nagarale R, Gohil G, Shahi V(2006) Recent developments on
ion-exchange membranes and electro-membrane processes. Adv
Colloid Interface Sci 119 : 97 130
154. Nawaz H, Shi J, Mittal G, Kakuda Y (2006) Extraction of
polyphenols from grape seeds and concentration by ultrafiltration. Sep Purif Technol 48(2):176181
155. Nigmet U, Yilmaz L, Yetis U (2009) Microfiltration/ultrafiltration as pretreatment for reclamation of rinsing waters of indigo
dyeing. Desalination 240:198208
156. Nong-xue Q, Yu-xia T, Shu-tao Q, Deng H (2009) Apple pectin
behavior separated by ultrafiltration. ASC 8(10):11931202
157. Nour-Eddine E, Ghidouche S, Ducrot P (2007) Flavonoids:
hemisynthesis, reactivity, characterization and free radical
scavenging activity. Rev Mol 12(9):22282258
158. Nuang Sim L, Yun Ye V, Chen A, Fane G (2011) Investigations
of the coupled effect of cake-enhanced osmotic pressure and
colloidal fouling in RO using cross-flow sampler-modified
fouling index ultrafiltration. Desalination 273:184196

Food Eng Rev (2011) 3:136158


159. Ogunbiyi O, Miles N, Hilal N (2008) The effects of performance
and cleaning cycles of new tubular ceramic microfiltration
membrane fouled with a model yeast suspension. Desalination
220:273289
160. OkamotoY KazushigeO, Glatz CE (2001) Harvest time effects
on membrane cake resistance of Escherichia coli broth. J Memb
Sci 190:93106
161. Onsekizogluv P, Bahceci KS, Acar MJ (2010) Clarification and
the concentration of apple juice using membrane processes: a
comparative quality assessment. J Memb Sci 352(12):160165
162. Ozoglu B (2002) Inhibition of enzymic browning in cloudy
apple juice with selected antibrowning agents. Food Control
13:213222
163. Pagan A, Conde J, Ibarz A, Pagan J (2010) Albedo hydrolysis
modelling and digestion with reused effluents in the enzymatic
peeling process of grapefruits. J Sci Food Agric 90(14):2433
2439
164. Pap N, Mahosenaho M, Pongracz E, Mikkonen H, Jaakkola M,
Virtanen V, Myllykoski L, Horvath-Hovorka Z, Hodur C, Vatai
G (2011) Effect of ultrafiltration on anthocyanin and flavonol
content of black currant juice (Ribes nigrum L.). Food Bioprocess Tech 1(4). doi:10.1007/s11947-010-0371-z
165. Perez S, Rodrguez-Carvajal M, Doco T (2003) A complex plant
cell wall polysaccharide: rhamnogalacturonan II. A structure in
quest of a function. Biochimie 85:109121
166. Petrotos KB, Lazarides HN (2001) Osmotic concentration of
liquid foods. J Food Process Eng 49(23):201206
167. Pinelo M, Zeuner B, Meyer A (2010) Juice clarification by
protease and pectinase treatments indicates new roles of pectin
and protein in cherry juice turbidity. FBP 88(23):259265
168. Pongsuriya K, Tatsuo K, Noriyuki I, Mitsuya S (2006) Changes
in the odours of apple juice during enzymatic browning. Food
Qual Prefer 17:497504
169. Queiroza C, Mendes Sopesa M, Fialhoa E, Valente-Mesquita V
(2008) Polyphenol oxidase: characteristics and mechanisms of
browning control. Food Rev Int 24:361375
170. Quevedo R, Jaramillo M, Daz O, Pedreschi F, Aguilera J (2009)
Quantification of enzymatic browning in apple slices applying
the fractal texture Fourier image. J Food Eng 95:285290
171. Rai P, Majumdar G, Dasgupta S, Deb S (2005) Understanding
ultrafiltration performance with mosambi juice in an unstirred
batch cell. J Food Eng 28(2):166180
172. Rai P, Majumdar GC, DasGupta S, De S (2005) Quantification
of flux decline of depectinized mosambi juice using unstirred
batch ultrafiltration. J Food Process Eng 28:359377
173. Rai P, Raia C, Majumdara G, DasGupta S, Deb S (2006)
Resistance in series model for ultrafiltration of mosambi (Citrus
sinensis (L.) Osbeck) juice in a stirred continuous mode.
J Memb Sci 283:116122
174. Rai P, Majumdar GC, DasGupta S, De S (2007) Modeling of
permeate flux of synthetic fruit juice and mosambi juice (Citrus
sinensis) in stirred continuous ultrafiltration. LWT 40:1765
1773
175. Ramadan A, Gyula V, Bekassy-Molnar E, Balint A (2005)
Investigation of ultra- and nanofiltration for utilization of whey
protein and lactose. J Food Process Eng 67:325332
176. Reid E, Liu X, Judd SJ (2008) Sludge characteristics and
membrane fouling in full-scale submerged membrane bioreactors. Desalination 219:240249
177. Richard-Forget FC, Goupy P, Nicolas J (1992) Cysteine as an
inhibitor of enzymic browning. 2. Kinetic studies. J Agric Food
Chem 40(11):21082113
178. Roorda JH, Van der Graaf JM (2001) New parameter for
monitoring fouling during ultrafiltration of WWTP effluent.
Water Sci Technol 43(10):241248

157
179. Sa IS, Matta VM, Cabral LM (2003) Concentration of pineapple
juice by membrane separation processes. Braz J Food Technol
6:5362
180. Sablani S, Goosen M, Al-Belushi R, Wilf M (2001) Concentration polarization in ultrafiltration and reverse osmosis: a
critical review. Desalination 141(3):269289
181. Santon M, Treiche H, Valduga E, Cabra L, Di Luccio M (2008)
Evaluation of enzymatic treatment of peach juice using response
surface methodology. J Sci Food Agric 88:507512
182. Saxena R, Gupta R, Saxena S, Gulati R (2001) Role of fungal
enzymes in food processing. Applied mycology and biotechnology. Agri Food Prod 1:353386
183. Saxena A, Tripathi B, Kumar M, Shahi V (2009) Membranebased techniques for the separation and purification of proteins:
an overview. COCIS 145(12):122
184. Schafer AI, Andritsos N, Karabelas AJ, Hoek EM, Schneider R,
Nystrom M (2004) Fouling in nanofiltration. In: Schafer AI,
Waite TD, Fane AG (eds) Nanofiltrationprinciples and
applications, vol 20. Elsevier, Oxford, pp 169239
185. Schuster E, Dunn-Coleman N, Frisvad J, van Dijck P (2002) On
the safety of Aspergillus nigera review. Appl Microbiol
Biotechnol 59(45):426435
186. Shahnawaz M, Shiekh A (2011) Analysis of viscosity of jamun
fruit juice, squash and jam at different compositions to ensure
the suitability of processing applications. IJPPB 3(5):8994
187. Shengmin L, Yaguang L, Turner E, Feng H (2007) Efficacy of
sodium chlorite as an inhibitor of enzymatic browning in apple
slices. Food Chem 104(2):824829
188. Shuying C, Darren L, Oatley P, Williams M, Wright CJ (2011)
Positively charged nanofiltration membranes: review of current
fabrication methods and introduction of a novel approach. Adv
Colloid Interface Sci 164(12):1220
189. Silva CM, Reeve DW, Husain H, Rabie H, Woodhouse K (2000)
Model for flux prediction in high-shear microfiltration systems.
J Memb Sci 173:8798
190. Smith KE, Bradley RL (1987) Efficacy of sanitizers using
unsoiled spiral-wound polysulfone ultrafiltration membranes.
J Food Prot 50(7):567572
191. Son S, Moon K, Lee C (2000) Kinetic study of oxalic acid
inhibition on enzymatic browning. J Agric Food Chem
48:20712074
192. Song L, Chen L, Ong S, Ng W (2004) A new normalization
method for determination of colloidal fouling potential in
membrane processes. J Colloid Interface Sci 271:426433
193. Srimanta P, Bharihoke R, Chakraborty S, Kumar S, Sirshendu
D, Sunando D (2008) An experimental and theoretical analysis
of turbulence promoter assisted ultrafiltration of synthetic fruit
juice. Sep Purif Tech 62:659667
194. Stopka J, Schlosser S, Domeny Z, Smogrovicov D (2000) Flux
decline in microfiltration of beer and related solutions of model
foulants through ceramic membranes. Pol J Environ Stud
9(1):6569
195. Sulaiman MZ, Sulaiman NM, Abdellah B (2001) Prediction of
dynamic permeate flux during cross-flow ultrafiltration of
polyethylene glycol using concentration polarization-gel layer
model. J Memb Sci 189:151165
196. Sutherland K (2008) Water filtration: bulk water filtration
techniques. Filtr Sep 45(10):1719
197. Taniguchi M, Kilduff JE, Belfort G (2003) Low fouling synthetic membranes by UV-assisted graft polymerization: monomer selection to mitigate fouling by natural organic matter.
J Memb Sci 222:5970
198. Tasselli F, Cassano A, Drioli E (2007) Ultrafiltration of kiwifruit
juice using modified poly (ether ether ketone) hollow fibre
membranes. Sep Purif Tech 57:94102

123

158
199. Tochi B, Wang Z, Ying Xu S, Zhang W (2009) Effect of stem
bromelain on the browning of apple juice. Am J Food Tech
4(4):146153
200. Torras C, Nabarlatz D, Montane D, Garcia-Valls R (2006)
Enzymatic membrane reactors based on polysulfone/activated
carbon. Desalination 199:438440
201. Torras C, Pallares J, Garcia-Valls R, Jaffin M (2009) Numerical
simulation of the flow in a rotating disk filtration module
235(13):122138
202. Tran-Ha M, Wiley D (1998) The relationship between membrane cleaning efficiency and water quality. J Memb Sci
145:99110
203. Tzahi Y, Cath V, Adams D, Amy Childress E (2004) Experimental study of desalination using direct contact membrane
distillation: a new approach to flux enhancement. J Memb Sci
228:516
204. Vaillant F, Jeanton E, Dornier M, OBrien GM, Reynes M,
Decloux M (2001) Concentration of passion fruit juice on an
industrial pilot scale using osmotic evaporation. J Food Eng
47(3):195202
205. Van der Bruggen B, Vandecasteele C (2002) Distillation vs.
membrane filtration: overview of process evolutions in seawater
desalination. Desalination 143:207218
206. Van der Bruggen B, Braeken L, Vandecasteele C (2002) Flux
decline in nanofiltration due to adsorption of organic compounds. Sep Purif Tech 29:2331
207. Vladisavljevic GT, Vukosavljevic P, Bukvic B (2003) Permeate
flux and fouling resistance in ultrafiltration of depectinized apple
juice using ceramic membranes. J Food Eng 60:241247
208. Vourch M, Balannec B, Chaufer B, Dorange G (2005) Nanofiltration and reverse osmosis of model process waters from the
dairy industry to produce water for reuse. Desalination
172:245256
209. Wagner J (2001) Membrane filtration handbook. Practical tips
and hints. Revision 2, 2nd edn. Osmonics Inc., Minnetonka
210. Wahab MA, Noraaini A (2002) Understanding the steric and
charge contributions in NF membranes using increasing MWCO
polyamide membranes. Desalination 147(13):205212
211. Wakeman RJ, Williams CJ (2002) Additional techniques to
improve microfiltration. Sep Purif Techn 26:318

123

Food Eng Rev (2011) 3:136158


212. Wang BJ, Wei TC, Yu ZR (2005) Effect of operating temperature on component distribution of West Indian cherry juice in a
microfiltration system. LWT 38:683689
213. Warczok J, Ferrando M, Lopez F, Guell C (2004) Concentration
of apple and pear juices by nanofiltration at low pressures.
J Food Eng 63:6370
214. Whitaker J, Lee CY (1995) Recent advances in chemistry of
enzymatic browning an overview: enzymatic browning and its
prevention in enzymatic browning and its prevention ACS
symposium series. American Chemical Society, Washington,
pp 27 Chapter 1
215. Yasan H, Zhijuan J, Shunxin L (2007) Effective clarification of
apple juice using membrane filtration without enzyme and
pasteurization pretreatment. Sep Purif Techn 57:366373
216. Yazdanshenas M, Reza Tabatabaee-Nezhad SA, Soltanieh M,
Roostaazad R, Khoshfetrat AB (2010) Contribution of fouling
and gel polarization during ultrafiltration of raw apple juice at
industrial scale. Desalination 258:194200
217. Ying PL, Abdul WM (2010) Effect of solution chemistry on flux
decline during high concentration protein ultrafiltration through
a hydrophilic membrana. Chem Eng J 159(13):9197
218. Yu J, Lencki RW (2004) Effect of enzyme treatments on the
fouling behavior of apple juice during microfiltration. J Food
Eng 63:413423
219. Zapata-Montoya JE, Guadix EM, Guadix A (2006) Long-term
effects of chemical cleaning in the performance of ultrafiltration
ceramic membranes. Desalination 200:316318
220. Zarate-Rodrguez E, Ortega-Rivas E, Barbosa-Canovas G
(2007) Effect of membrane pore size on quality of ultrafiltered
apple juice. Food Chem 104:824829
221. Zer-Ran Y, Chien-Ching H, Yih-Ming W, Chun-Li S, Be-Jen W
(2007) Physiochemical, antioxidant and whitening properties of
extract from root cortices of mulberry as affected by membrane
process. LWT 40(5):900907
222. Zhong J, Xiaojuan S, Wang C (2003) Treatment of oily wastewater produced from refinery process using flocculation and
ceramic membrane filtration. Sep Purif Tech 32:9398
223. Zoha KD, Stenstrom MK (2002) Application of a membrane
bioreactor for treating explosives process wastewater. Water Res
36(4):10181024

Vous aimerez peut-être aussi