Sie sind auf Seite 1von 10

Journal of Food Engineering 187 (2016) 14e23

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Determination of fouling mechanisms in polymeric ultraltration


membranes using residual brines from table olive storage wastewaters
as feed
Carlos Carbonell-Alcaina a, *, Mara-Jose
 Corbato
 n-Ba 
guena a, Silvia Alvarez-Blanco a, b
,
Mara Amparo Bes-Pia  a,b
 Antonio Mendoza-Roca
, Jose a,b
, Laura Pastor-Alcan~ iz c

a
Research Institute for Industrial, Radiophysical and Environmental Safety (ISIRYM), Universitat Polit
ecnica de Val
encia, C/Camino de Vera s/n, 46022
Valencia, Spain
b
Department of Chemical and Nuclear Engineering, Universitat Polit ecnica de Val
encia, C/Camino de Vera s/n, 46022 Valencia, Spain
c
Depuracion de Aguas del Mediterra neo (DAM), Avda. Benjamin Franklin, 21, Parque Tecnologico, 46980, Paterna Valencia, Spain

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

Article history: In this work, the fouling mechanisms that dominate the ultraltration of residual brines from table olive
Received 30 November 2015 packing plant wastewaters were investigated. For that purpose, Hermia's models adapted to crossow
Received in revised form ltration, resistance-in-series model and a model combining intermediate blocking and cake formation
19 April 2016
mechanisms were tted to the experimental data. Tests were performed with a 5 kDa polyethersulfone
Accepted 23 April 2016
membrane at transmembrane pressures between 1 and 3 bar and crossow velocities between 2.2 and
Available online 26 April 2016
3.7 m s1. Results demonstrated that the resistance-in-series model was the most accurate to predict
permeate ux evolution with time. The predominant fouling mechanism was cake formation followed by
Keywords:
Ultraltration
intermediate blocking/adsorption. The fouling resistances that were determined by means of the resis-
Polyethersulfone membranes tance in series model were tested using a well-established mathematical model proposed by Mondal and
Table olive packing plant wastewaters De that also combines both fouling phenomena (intermediate pore blocking and cake formation). Results
Mathematical models demonstrated that the predicted resistances are consistent with those determined by Mondal and De's
Fouling mechanisms model.
2016 Elsevier Ltd. All rights reserved.

1. Introduction particles from liquid streams (Wang and Song, 1999; Barredo-
Damas et al., 2012). In this work, UF is considered to remove sus-
Spain is the main producer of table olives with 57,3,500 t year1, pended particles and macromolecules from residual brines from
which is around 22.1% of the total world production (International table olive production plants. A subsequent nanoltration (NF) step
Olive Oil Council, 2014). During the production process, large could be performed to recover the phenolic compounds.
amounts of water are used, generating high volumes of wastewater. Currently there are many studies focused on membrane treat-
Three types of wastewater are obtained: debittering, washing and ment of wastewaters from olive oil production, such as, olive mill
fermentation brines (Benitez et al., 2003). Fermentation brine wastewater (OMW) residue from a three-phase production method
wastewater is characterized by an acidic pH (around 4), a high and alperujo residue from a two-phase production method. These
conductivity (80e115 mS cm1), a high concentration of total sus- residues have a high chemical oxygen demand and a high phenolic
pended solids (0.2e2 g L1), dissolved chemical oxygen demand compound concentration, but unlike fermentation brine, its con-
(COD: 10e35 g O2$L1) and phenolic compounds (4.0e6.0 g of ductivity is much lower (between 4.00 and 13.98 mS cm1 for
tannic acid L1) (Garrido Fern
andez et al., 1997). Ultraltration (UF) OMW and 0.88e4.76 mS cm1 for alperujo) (Paredes et al., 1999;
is one of the most used techniques in industry to: concentrate, Alburquerque et al., 2004). Nanoltration (NF) processes have
separate or purify macromolecules, colloids and suspended been considered by most of the authors to recover and concentrate
high added value compounds from olive oil production wastewa-
ters. In order to improve the performance of the NF process, a pre-
* Corresponding author.
treatment with UF has also been proposed (Galanakis et al., 2010;
E-mail address: carcaral@upvnet.upv.es (C. Carbonell-Alcaina). Paraskeva et al., 2007). However, the number of studies on the

http://dx.doi.org/10.1016/j.jfoodeng.2016.04.016
0260-8774/ 2016 Elsevier Ltd. All rights reserved.
C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23 15

treatment of fermentation brine wastewater by membrane tech- 2014).


nology is very limited. Nowadays, the technologies that have shown In the same way as Ho and Zydney, some other authors com-
more favorable results for recovering desirable compounds from bined two different fouling mechanisms providing a model with
the brines have been membrane processes. This can be combined strong theoretical background. For instance, Mondal and De pub-
with adsorption processes using active carbon or ion exchange lished two different articles which describe in detail a generalized
resins (Bo dalo et al., 2008). The rejection of total suspended solids, model for steady state continuous ltration (Mondal and De, 2009,
dissolved COD and phenolic compounds in UF process treating olive 2010). The proposed model resulted from the combination of two
mill wastewater (OMW) and table olive wastewaters was investi- different fouling mechanisms: on their rst article, Mondal and De
gated by El-Abbassi (El-Abbassi et al., 2014). took into account the resistance due to complete pore blocking and
Membrane fouling is one of the major problems of UF processes, that related to the formation of a cake on the membrane surface
reducing the permeate ux and decreasing its economic and (Mondal and De, 2009); while the second article combined the
technological viability (Cheryan and Alvarez, 1995). During an UF resistances due to intermediate pore blocking and cake formation
process where fouling occurs, the initial permeate ux shows a mechanisms (Mondal and De, 2010). In addition, Bolton et al.
sharp decline which is followed by a long and gradual ux decline developed ve new models based on the four classical fouling
over time (Field et al., 1995). Therefore, the study of the evolution of mechanisms (standard blocking, complete and intermediate
permeate ux over time during UF is an important factor to be blocking and cake formation mechanisms) and the Darcy's law for
considered when selecting the optimum operating conditions both constant pressure and constant ow operation modes (Bolton
(Vincent-Vela et al., 2010). In order to predict the UF membrane et al., 2006). Their theoretical hypotheses resulted in general
fouling and its performance, mathematical models have been equations with two tting parameters that were much simpler
developed by several authors. In literature, empirical and theoret- than those taken as references by the authors. Their results also
ical models that describe the permeate ux decline, with time in demonstrated that the new models predicted with good accuracy
UF, can be found. Empirical models provide high precision, but they the experimental data. Using the equations developed by Bolton
cannot satisfactorily explain the fouling mechanisms involved in et al., other authors, such as Rezaei et al., tted those general
membrane ltration. Theoretical models can help to better under- equations for the combined models to the experimental data ob-
stand the phenomenon of fouling, but if experimental data is not tained during whey crossow microltration (Rezaei et al., 2011).
used to estimate some of the parameters their predictions are not They also compared the tting accuracies of the combined models
very precise. Thus, semi-empirical models whose parameters have with the classical ones that account only for one fouling mechanism
a physical meaning are usually preferred to explain the fouling at a time. Their results demonstrated that the combined models
phenomena that takes place in membrane processes and to achieve were able to better predict the permeate ux decline at low time
an accurate prediction of permeate ux decline (Vincent Vela et al., scales, but classical models provided higher accuracies at those
2009; Mah et al., 2012). Depending on the fouling mechanism, four experimental conditions where cake resistance was the predomi-
situations may be described: (a) if the solute particle size is higher nant fouling mechanism.
than the membrane pore; particles are deposited on the surface of Although there are several mathematical models available in the
the membrane blocking the entrances of the pores completely; (b) literature to determine fouling mechanisms, Hermia's models and
if solute particle and membrane pores size are similar, some their adaptations to crossow ltration as well as resistance-in-
membrane pores can be partially blocked; (c) if solute particle size series models are the most accepted. These have also been used
is smaller than membrane pores, inside pores of membrane can be to predict permeate ux decline, with time, by other authors
blocked and irreversible fouling may appear; (d) sometimes the (Carrere et al., 2001; Turano et al., 2002; Vincent Vela et al., 2009).
fouling layer deposited on the membrane surface may form a cake Vincent Vela et al. (2009) tted the Hermia's models adapted to
layer (Ruby Figueroa et al., 2011; Corbato n-Baguena et al., 2013). crossow to the experimental data of permeate ux versus time
Among the different theoretical models available in the litera- obtained during the UF of polyethylene glycol solutions. They
ture to determine fouling mechanisms, one of the most widely used demonstrated that at high transmembrane pressures and low
is the one developed by Ho and Zydney (Ho and Zydney, 2000). The crossow velocities, the intermediate blocking was the predomi-
general equation of this model accounts for the combination of nant fouling mechanism. Corbato n-Baguena et al. (2015a) tted
pore blockage and cake formation without time division of the different models to the experimental data obtained during the UF
permeate ux curve. The authors tted the model to the BSA of whey model solutions. They reported that the combination of
microltration experimental data. Based on the Ho and Zydney's complete blocking and cake formation mechanisms resulted in
model, recent works have used their mathematical assumptions to more accurate predictions of the permeate ux decline. Carre re
t the experimental data of different UF processes and also, modify et al. (2001) proposed a resistance-in-series model that consid-
the original model (Muthukumaran et al., 2005; Peng and ered the membrane resistance, the cake resistance and the
Tremblay, 2008; Karasu et al., 2010; Corbato  n-Ba
guena et al., adsorption and concentration polarization resistance. This was
2013; Liu et al., 2014; Tien et al., 2014; Astaree et al., 2015). For done to estimate permeate ux decline in the crossow micro-
instance, Astaree et al. studied membrane fouling mechanisms ltration of lactic acid fermentation broths. They reported that the
caused by BSA, dextran and humic acid solutions by tting the adsorption and concentration polarization resistances dominated
model proposed by Ho and Zydney and modifying it to consider the the ltration process. Turano et al. (2002) also interpreted the
hydrophilic nature of the membranes used and the pre-ltration experimental data of permeate ux variation with the time ob-
effect of the foulant deposit layer. With these two new factors, tained during the UF of olive mill wastewaters by using a
they demonstrated that better agreements were obtained in com- resistance-in-series model. They indicated that higher values of
parison to the original mathematical model (Astaree et al., 2015). turbulence and thus higher crossow velocities resulted in lower
On the other hand, Tien et al. used the experimental data presented values of specic cake resistance.
by Ho and Zydney to validate their new rational model, which was In this work, the effect of transmembrane pressure (TMP) and
based on the deep bed ltration theories and the equations for crossow velocity (CFV) on ux decline when UF was used to treat
particle retention within membrane media. Their model was able to residual brine from a table olives packing plant (TOPP) was inves-
predict the experimental data provided by Ho and Zydney and its tigated. The samples were previously ltered through a 60 mm
general equation was simpler than the original one (Tien et al., cartridge lter. In order to understand the predominant fouling
16 C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23

mechanisms affecting permeate ux decline when UF membranes blocking model, which considers that foulant molecules cannot
were fouled with the residual brine, several mathematical models penetrate inside the porous structure and are deposited on previ-
were taken into account: Hermia's models adapted to crossow ously settled molecules. The complete blocking model (n 2), oc-
ltration, a resistance-in-series model and a model that combines curs when a solute molecule completely seals a pore entrance
the complete blocking and cake formation fouling mechanisms without penetrating inside the membrane pores and solute mole-
developed by Hermia. It is important to highlight the practical cules form a monomolecular layer on the membrane surface.
relevance of this type of empirical models based on exponential Finally, solute molecules that have a smaller size than the mem-
equations, since recent works were published at this regard. For brane pores can pass through the porous structure and block the
instance, Lin et al. used an exponential model with four tting inner pore walls. This corresponds to the standard blocking model
parameters to predict the entire permeate ux curve obtained in (n 1.5).
the ultraltration of protein aqueous solutions (Lin et al., 2008).
They divided the decrease in permeate ux according to two
2.2. Combined model (intermediate pore blocking and cake
fouling phenomena: intermediate blocking for the rst minutes of
formation)
operation and cake formation fouling for the rest of the ultral-
tration curve. Yee et al. studied the crossow ultraltration of whey
According to the literature (Field et al., 1995; Ho and Zydney,
by tting an exponential equation to the experimental data ob-
2000), the typical evolution of permeate ux during the UF time
tained in the fouling experiments (Yee et al., 2009). The empirical
can be divided in two stages: a great decline in permeate ux
model was compared to Ho and Zydney's one, and tting results
during the rst minutes of operation because of a pore blocking
demonstrated that the theoretical Ho and Zydney's model did not
phenomenon and a gradually slow ux decline caused by the
provide high accuracies at great time scales. However, model t-
accumulation of foulant molecules on the membrane surface and
tings using the exponential pattern were in a good agreement with
the formation of a cake layer on it. Along with this evolution, de la
the experimental data. Corbato n-Baguena et al. used empirical
Casa et al. (2008) developed a combined model taking into account
models with theoretical background to t the experimental data
the classical dead-end Hermia's equations for the complete block-
obtained during the ultraltration of enzymatic solutions
ing and cake formation model. In this work, a similar combination
(Corbato  n-Ba
guena et al., 2015b). These models were based on
of the intermediate pore blocking and cake formation models
exponential decay patterns whose main parameters have physical
described by Hermia was adapted to crossow operating mode (Eq.
meaning (for instance, ratio among original membrane character-
(2)):
istics and those achieved after fouling and resistances due to
complete blocking and cake formation). In the literature, most of Jcombined model a,Jintermediate blocking model
the mathematical models were tested with model solutions instead
of real solutions. Moreover, the fouling mechanism that causes ux 1  a,Jcake layer formation model (2)
reduction when table olive residual brine is ultraltered has not
been investigated so far. In addition, Astudillo-Castro explained the where a is the fraction of blocked membrane pores. In addition, Eq.
fundamentals of the exponential equations proposed in this work, (2) includes two different model constants (Ki for the intermediate
which are based on the direct relationship between the maximum blocking model and Kcf for the cake layer formation model) as they
resistance value (achieved at the steady state) and that obtained at are the constants that correspond to Hermia's equations adapted to
a certain time (Astudillo-Castro, 2015). These authors applied the crossow operation for the intermediate pore blocking and the
abovementioned mathematical equations to describe the limiting cake formation models, respectively. The physical meaning of these
permeate ux as a function of transmembrane pressure. Another two parameters is the following: Ki represents the membrane
novelty of this work corresponds to the validation of the tted surface blocked per unit of total permeate volume and unit of initial
parameters obtained using a well-established, theoretical model membrane surface porosity, while Kcf is the ratio between the
available in the literature (resistance in series model developed by characteristics of the cake layer on the membrane surface and those
Mondal and De, 2010). of the original, unfouled membrane (Vincent Vela et al., 2009).

2. Ultraltration modelling 2.3. Resistance-in-series model

2.1. Hermia's models adapted to crossow ltration The resistance-in-series model is based on the Darcy's law,
which relates the permeate ux (J) with the transmembrane
Based on the classical Hermia's models for constant pressure pressure (DP), the feed solution viscosity (m) and the total hydraulic
dead-end ltration (Hermia, 1982), several authors (Field et al., resistance (Rt) (Eq. (3)).
1995; de Barros et al., 2003; Vincent Vela et al., 2009) included
DP
the steady-state ux related to the back-transport mass transfer to J K,DP (3)
adapt the classical models to a crossow conguration. Eq. (1)
m,Rt
shows the general model equation for the adapted Hermia's This model describes the total hydraulic resistance during the
models: ltration process as the sum of several resistances: the original
membrane resistance (Rm), the adsorption and concentration po-
dJ
 KJ  Jss J 2n (1) larization resistance (Ra) and the cake layer resistance (Rcf) (Eq. (4))
dt re et al., 2001; Turano et al., 2002):
(Choi et al., 2000; Carre
where J is the permeate ux, Jss is the steady-state permeate ux, K Rt Rm Ra Rcf (4)
is the model constant and n is the model parameter related to the
four different fouling mechanisms. If the value of n is zero, a cake According to Carrere et al., the experimental variation of Rt, Ra
layer is formed on the membrane surface due to the accumulation and Rcf follows an exponential trend (Carrere et al., 2001, 2002). In
of solute molecules that are larger than membrane pores. When addition, both adsorption and concentration polarization phe-
n 1, the fouling mechanism corresponds to the intermediate nomena are simultaneously determined in crossow ltration
C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23 17

experiments. Thus, Ra can be expressed as an exponential function permeate ux for low time scales were determined. After that,
of the steady-state resistance (R'a), the rate of molecules deposition using the value of RIPB at time t1 and substituting the values of Rcf
on the membrane surface (b) and the ltration time (t), according to from the resistance-in-series model (described in Section 2.3) in Eq.
Eq. (5). Based on Eq. (5), in this work a general equation that ac- (11), the predicted values of permeate ux for time scales greater
count for Rcf as a function of time was newly proposed and repre- than t1 were determined. Finally, the permeate ux predicted
sented in Eq. (6), where R'cf is the steady-state cake resistance and c values were compared to the experimental data and to the resis-
is the rate of cake growth: tance in series method data. The goodness of the ts (in terms of R2
and SD) was determined.
Ra R0a 1  expb,t (5)
3. Materials and methods
Rcf R0cf 1  expc,t (6)
3.1. Feed samples
Substituting Eq. (5) and Eq. (6) in Eq. (4) and considering the
Darcy's law (Eq. (3)), the general equation for the resistance-in- This work was performed using different real solutions from
series model can be obtained (Eq. (7)): residual table olive brine, supplied by a TOPP. Due to the high
suspended solids content, a ltration step with a polyester car-
DP
J   (7) tridge lter of 60 mm pore was carried out. Four different samples
m$ Rm R0a 1  expb$t R0cf 1  expc$t were provided by the TOPP. The average characteristics of the
ltered samples are shown in Table 1. The samples were stored at
5  C.

2.4. Validation of resistance-in-series model 3.2. Analytical methods

The constants obtained in this work for the resistance-in-series Feed and permeate samples were characterized. The pH and the
model were validated using a model developed by Mondal and De, conductivity were measured with a pH-Meter GLP 21 and EC-
2010. This model is also based on the resistance in series mecha- Meter GLP 31 (Crison, Spain). Total suspended solids (TSS) were
nism and combines the resistances of two classical fouling mech- determined according UNE 77,034 by means of glass microbre
anisms (intermediate pore blocking and cake formation). According lters with 1.2 mm pore size, using samples of 25 mL and differing
to this model, at low time scales (up to a certain time point, t1), weights before and after drying the microbre lters at 105  C, for
membrane fouling resistance is due to the intermediate pore 2 h. Turbidity was measured considering the UNE-EN ISO 7027
blocking phenomenon and as a result, permeate ux sharply de- standard method with a turbidimeter (D-112, DINKO, Spain). Dis-
clines with time (Eq. (8)). Therefore, permeate ux variation with solved COD was determined with Hach Lange kits, LCK014 (LCK,
time can be described as: Germany), dissolving the samples 10e12 times in order to remove
the interferences caused by the high chloride concentration.
DP
J (8)
m$Rm RIPB 3.3. Equipments and procedures

where RIPB is the intermediate pore blocking resistance. This


3.3.1. Ultraltration
resistance can be expressed as a function of time by means of the
The feed samples for the UF process were residual olive brine
intermediate pore blocking constant (KIPB), according to Eq. (9).
from a TOPP. This had been previously ltered using a 60 mm
Therefore, the general model equation up to t1 is Eq. (10).
polyester mesh cartridge lter (CA-0202-00, model GT). The auto-
DP mated UF laboratory plant regulated transmembrane pressure
RIPB $K $t (9) (TMP), cross ow velocity and temperature. A Rayow membrane
m IPB
module from Orelis (France) was utilized, congured to work with
only one membrane in cross-ow mode. A Microdyn Nadir (Ger-
DP
J   (10) many) polyethersulfone (PES) membrane with a MWCO of 5 kDa
m$ Rm DmP$KIPB $t (Nadir UP005) was tested. The total active surface of the membrane
was 0.0125 m2. The Rm value for the UP005 membrane used in the
For time scales from t1 to the end of the ltration process, the experiments was 8.9216$1012 0.8057$1012 m1, obtained from the
formation of a cake on the membrane surface becomes the pre- observed linear relationship between pure water permeate ux and
dominant fouling mechanism and therefore, the general model TMP (R2 0.9965 0.0030). Darcy's law (Eq. (3)) was used for the
equation considers the effect of cake formation from t1 to the end of calculation. This value was considered as a constant to t the
the ltration process as well as the effect of intermediate pore resistance-in-series model to the experimental data.
blocking at t1. Then permeate ux decline with time is described by
Eq. (11): Table 1
Characteristics of feed ltered with a 60 mm polyester cartridge lter (mean values
DP for all the feed samples).
J   (11)
m$ Rm RIPB t1 Rcf t  t1 Parameter Mean value Standard deviation

pH 4.0 0.2
Therefore, in order to validate the constants obtained for the Conductivity (mS/cm) 85.3 12.1
resistance-in-series model, Eq. (10) was rstly tted to the exper- Turbidity (NTU) 262.8 92.8
imental data up to the selected t1 values for each experimental TSS (mg/L) 411.7 80.0
condition tested. Once the values of KIPB were obtained, the resis- Dissolved COD (mg/L) 16547.7 3902.0
Phenolic compounds (mg/L Tyrosol eq) 1100.9 325.7
tance RIPB was calculated using Eq. (9) and the predicted values of
18 C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23

Fouling and cleaning tests were carried out with varying TMP 3.5
and CFV, set at 1e3 bar and 2.2e3.7 m s1 respectively, for 2.5 h.

Permeate flux 10 5 (m3m-2s-1)


During this time temperature was kept at a constant 25  C. After 3.0
this period, steady state was achieved, allowing permeate and 2.2 m/s (Sample S2)
retentate to be recycled back to the feed tank. Permeate ux was 2.5 2.9 m/s (Sample S4)
monitored with a precision balance from Kern (Germany) and the 3.7 m/s (Sample S3)
2.0
collected data was recorded with a data acquisition system. After Resistance in Series Model
each run, it was checked that if was restored to the initial perme- 1.5
ability. A permeability recovery higher than 95% was required
before further testing could be carried out. Membrane cleaning was 1.0
performed by rinsing with osmotic water at CFV of 2.2 m s1 and
TMP of 0.6 bar during 9 min. If the initial permeability was not 0.5
restored membrane was cleaned with chemicals using a concen-
trated basic solution of NaOH (pH 11) (Panreac, Spain) and acid 0.0
solution in water of citric acid (1% w/v) (Panreac, Spain) at CFV of 0.0 50.0 100.0 150.0 200.0
2.2 m s1 and TMP of 0.6 bar during 5 min, respectively. After each Time (min)
chemical cleaning, the system was rinsed with osmotic water at the Fig. 2. Evolution of permeate ux with time during the UF experiments at TMP of 2 bar
same operating conditions for other 9 min. and model tting to the experimental data.

3.4. Model tting


4.5
The model was tted to the experimental data using MathCad

Permeate flux 10 5 (m3m-2s-1)


4.0
15 (PTC Needham, EE.UU). This was carried out by means of the
MathCad Gent algorithm, which uses an optimized version of the 3.5 2.2 m/s (Sample S3)
Levenberg-Marquadt method, minimizing the difference between 3.0 2.9 m/s (Sample S3)
the predicted and experimental results. The quality of the tting for 3.7 m/s (Sample S3)
2.5
each operating condition tested was evaluated in terms of the Resistance in Series Model
regression coefcient (R2) and the standard deviation (SD). 2.0
1.5
4. Results and discussion
1.0
Figs. 1 to 3 show the variation of permeate ux with time for the 0.5
combinations of TMP and CFV tested. As it can be observed, the
0.0
values of steady state permeate ux were low for all the operating
conditions tested because of fouling. At a TMP of 1 bar, the variation
0.0 50.0 100.0 150.0 200.0
Time (min)
of CFV within the range considered in this work did not produce
signicant variations in permeate ux. However, at 2 and 3 bar ux Fig. 3. Evolution of permeate ux with time during the UF experiments at TMP of 3 bar
variation with TMP at certain CFV was lower than ux variation and model tting to the experiment data.
with CFV at certain TMP. Therefore, these tested pressures were in
the range where ux reaches the critical ux, as it can be observed
from Fig. 4. In this gure the stationary permeate ux is plotted 3.5
Permeate flux 10 5 (m3m-2s-1)

against TMP for the different CFV tested. It is observed that the
stationary permeate uxes are far from the osmotic water line. The 3.0
Osmoc water
stationary permeate ux increased with TMP up to a point at which CFV: 2.2 m/s
permeate ux remains constant or decreases with TMP. In this 2.5
CFV: 2.9 m/s
CFV: 3.7 m/s
2.0
2.5
1.5
Permeate flux 10 5 (m3m-2s-1)

2.0 1.0
2.2 m/s (Sample S2)

1.5 2.9 m/s (Sample S1) 0.5


Resistance in Series Model
0.0
1.0 0.0 1.0 2.0 3.0 4.0
TMP (bar)
0.5 Fig. 4. Steady state permeate ux versus transmembrane pressure (TMP) for the
different crossow velocities (CFV) tested.

0.0
0.0 50.0 100.0 150.0 200.0 region, if pressure increases, an increase in particle concentration
Time (min)
on the membrane boundary layer takes place (Miller et al., 2014).
Fig. 1. Evolution of permeate ux with time during the UF experiments at TMP of 1 bar The presence of a high concentration of solutes in table olive re-
and model tting to the experimental data. sidual brine has a great effect on their concentration in the mass
C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23 19

transfer boundary layer as well as on the membrane surface, was.


affecting the development of the gel layer. Therefore, the effect of The effect of turbidity on ux decline is also observed in Table 2.
TMP and CFV on concentration polarization is signicant (Mondal This table shows that, for the same operating conditions, ux
et al., 2011). Thus, as it can be observed in Figs. 1e3, permeate decline increased when turbidity increased (for example when
ux decline decreases with CFV and increases with TMP. samples S2 and S4 are treated at 2 bar and 2.9 m s1). It ca be also
observed that ux decline was greater for sample S1 at 1 bar and
2.9 m s1 than that for sample S4 at 2 bar and 2.9 m s1. This is due
4.1. Modelling results to the effect of turbidity, which caused higher ux reduction even
at lower transmembrane pressures. Sample S4 was much more
The evolution of permeate ux with time during the UF exper- turbid than sample S1. This effect was not observed however when
iments for each transmembrane pressure, crossow velocity and ux decline for sample S1 at 1 bar and 2.9 m s1 is compared to ux
feed solution tested is shown in Figs. 1 to 3. As transmembrane decline for sample S2 at 2 bar and 2.9 m s1, because the difference
pressure increased, ux decline during the rst minutes of opera- between their turbidity was much lower.
tion was more rapid, as it was expected (Ho and Zydney, 2000). This Table 3 shows the tting accuracy results for all the models
behaviour corresponded to the existence of pore blocking phe- tested in terms of the regression coefcient (R2) and standard de-
nomena, which are mostly responsible for membrane fouling at viation (SD). The values of R2 and SD for the model with the highest
low time scales. In addition, the higher the crossow velocity was, tting accuracy are highlighted in bold in the table. It is important
the higher the steady-state permeate ux was. According to other to highlight that Hermia's standard blocking model adapted to
authors (Lin et al., 2004), the shear stress and turbulence produced crossow did not t to the experimental data. This model considers
by an increase in crossow velocity may prevent membrane surface that solute molecules are smaller than membrane pore size and
from severe fouling. Therefore, the permeate ux achieved at the thus they can penetrate inside the membrane porous structure.
end of the UF run increased as CFV increased. Therefore, it can be concluded that internal pore blocking is not
The percentage of ux decline reached at the end of the UF signicant. As it can be observed, the predictions of the resistance-
experiments for each feed solution and operating conditions tested in-series and combined models were the most accurate for all the
is shown in Table 2. As it can be observed, permeate ux decline transmembrane pressures, crossow velocities and feed solutions
signicantly increased as the transmembrane pressure applied tested. At 2 and 3 bar, the accuracy of both models was similar.
increased from 1 to 3 bar. For instance, for the sample S2, ux However, at 1 bar, the accuracy of the combined model was worse
decline increased from 69.02 to 74.91% when the transmembrane than the accuracy of the resistance-in-series model. Values of R2
pressure increased from 1 to 2 bar at a given crossow velocity and SD for the resistance-in-series model ranged from 0.953 to
(2.2 m s1). In addition, for the sample S3, at 3.7 m s1, permeate 0.996 and 0.040 to 0.016, respectively. Figs. 1 to 3 show the tting of
ux decline increased from 65.77 to 76.40% from 2 to 3 bar. The the resistance-in-series model to the experimental data. Table 3
initial permeate ux decline rate increases with transmembrane also allows comparison between Hermia's models and the com-
pressure due to the higher initial ux generated by TMP, which bined model. As it was explained in Section 2.2., the combined
allows the passage of more water through the membrane, before model considers the intermediate pore blocking and the cake for-
the concentration polarization phenomenon was entirely devel- mation mechanisms proposed by Hermia in order to predict the
oped (Chen and Kim, 2006). Moreover, as TMP increases there is a decline of permeate ux with time. In this case, as shown in Table 3,
greater transport of solute molecules towards the membrane sur- this combination of mechanisms resulted in a greater tting ac-
face due to the increase in the driven force. Thus, a greater decrease curacy than that obtained with the individual Hermia's models
in permeate ux was observed at high transmembrane pressures. adapted to crossow.
On the other hand, the trend observed as the crossow velocity The values of model parameters for the best tting model (the
increased was the opposite as that reported for the transmembrane resistance-in-series one) are shown in Table 4. As it was described
pressure, as expected, since permeate ux decline decreased with in Section 2.3, a high total hydraulic resistance causes great
crossow velocity for a xed value of transmembrane pressure. decrease in permeate ux. The total hydraulic resistance is the sum
Crossow velocity has a favorable inuence on membrane ltration of the original membrane resistance (Rm), the adsorption and
performance: the shear stress caused at high crossow velocities concentration polarization resistance (Ra) and the cake layer
may reduce concentration polarization and solute precipitation, resistance (Rcf), according to Eq. (4). While Rm is characteristic of the
preventing the membrane from a severe fouling (Lin et al., 2004). membrane and remains constant during the test, Ra and Rcf may
Therefore, the greatest the crossow velocity was for a given vary with time. The values of Ra and Rcf shown in the table corre-
transmembrane pressure, the lowest the permeate ux decline spond to the stationary values of the resistances. From Table 4, it
can be observed that Rcf is greater than Ra, therefore cake formation
Table 2 dominated permeate ux decline at the steady state for the oper-
Percentage of ux decline at the end of the UF experiments. ating conditions considered. Fig. 5 shows the evolution of the
different resistances with time for the sample S3 at 3 bar and
TMPa (bar) CFVb (m s1) Sample Feed turbidity (NTU) Flux decline (%)
3.7 m s1. It can be observed that at the beginning of the test pre-
1 2.2 S2 314.3 69.02 dominates the adsorption and concentration polarization resis-
2.9 S1 367.5 68.97
2 2.2 S2 314.3 74.91
tance. For all the tests performed with different samples and at
2.9 S2 314.3 72.39 different operating conditions the trend was similar. At the begin-
2.9 S4 183.4 67.26 ning of the run a rapid increase of the adsorption and concentration
3.7 S3 185.9 65.77 polarization resistance (Ra) takes place. However, the increase of
3.7 S4 183.4 64.65
the cake layer resistance (Rcf) with time is more gradual, but it
3 2.2 S3 185.9 84.22
2.2 S1 367.5 86.53 surpasses the value of Ra.
2.9 S3 185.9 79.98 As it was abovementioned, the initial permeate ux decline with
3.7 S3 185.9 76.40 time is due to the pore blocking phenomena. As it can be observed
a
TMP: transmembrane pressure. in Figs. 1e3, the higher the transmembrane pressure was, the more
b
CFV: crossow velocity. rapid the permeate ux decline was during the rst minutes of the
20 C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23

Table 3
Models tting accuracy: values of R2 and SD. Values in bold correspond to the best tting model.

TMPa (bar) CFVb (m s1) Sample Feed turbidity (NTU) Complete Intermediate Cake formation Combined Resistance-in-
blocking blocking model series

R2 SDc R2 SDc R2 SDc R2 SDc R2 SDc

1 2.2 S2 314.3 0.401 0.122 0.703 0.084 0.896 0.051 0.894 0.059 0.969 0.034
2.9 S1 367.5 0.525 0.108 0.760 0.076 0.908 0.050 0.850 0.068 0.953 0.040
2 2.2 S2 314.3 0.643 0.110 0.835 0.076 0.960 0.040 0.990 0.022 0.989 0.023
2.9 S2 314.3 0.527 0.103 0.747 0.076 0.893 0.050 0.984 0.022 0.985 0.022
2.9 S4 183.4 0.590 0.092 0.743 0.073 0.883 0.051 0.990 0.017 0.988 0.018
3.7 S3 185.9 0.494 0.109 0.676 0.088 0.852 0.062 0.992 0.016 0.986 0.020
3.7 S4 183.4 0.603 0.078 0.739 0.064 0.862 0.048 0.984 0.017 0.980 0.019
3 2.2 S3 185.9 0.713 0.140 0.889 0.094 0.986 0.038 0.995 0.023 0.996 0.023
2.2 S1 367.5 0.186 0.187 0.376 0.168 0.665 0.122 0.984 0.034 0.984 0.030
2.9 S3 185.9 0.670 0.120 0.836 0.089 0.961 0.048 0.996 0.017 0.994 0.020
3.7 S3 185.9 0.626 0.105 0.783 0.082 0.919 0.053 0.994 0.015 0.992 0.016
a
TMP: transmembrane pressure.
b
CFV: crossow velocity.
c
SD: standard deviation.

Table 4
Values of model parameters for the resistance-in-series model at the different operating conditions tested.

TMPa (bar) CFVb (m s1) Sample Feed turbidity (NTU) Model parameters

Rac (1013 m1) Rcfd (1013 m1)

1 2.2 S2 314.3 0.661 1.730


2.9 S1 367.5 0.669 1.557
2 2.2 S2 314.3 1.062 1.841
2.9 S2 314.3 1.249 1.454
2.9 S4 183.4 0.733 1.012
3.7 S3 185.9 0.818 0.867
3.7 S4 183.4 0.657 0.804
3 2.2 S3 185.9 1.614 2.759
2.2 S1 367.5 2.279 2.739
2.9 S3 185.9 1.259 1.975
3.7 S3 185.9 1.102 1.447
a
TMP: transmembrane pressure.
b
CFV: crossow velocity.
c
Ra: adsorption and concentration polarization resistance.
d
Rcf: cake layer resistance.

6.0 rst minutes of operation and results in a rapid ux decline; while


the second step is due to the formation and growth of a cake layer
on the membrane surface, which leads to a slow ux decline until a
Resistance 1013 (m-1)

5.0
Rt steady-state value is achieved. This pattern is related to the values
Rcf of Ra and Rcf. The parameter Ra is related to the pore blocking and
4.0
Ra absorption mechanism and the parameter Rcf is related to the cake
Rm formation. It is expected that, as transmembrane pressure in-
3.0
creases, the values of Ra also increase, due to the more severe
decrease in permeate ux observed in Fig. 3 compared to Fig. 1. This
2.0
trend is clearly observed in Table 4 for all the crossow velocities
tested when transmembrane pressure increased from 1 to 3 bar
1.0
independently of the feed sample. Regarding the effect of crossow
velocity on Ra, it can be observed that this parameter decreased as
0.0 crossow velocity increased. For example, the values of Ra for
0.0 50.0 100.0 150.0 200.0 sample S3 (turbidity of 185.9 NTU) at a transmembrane pressure of
Time (min) 3 bar decreased from 16.140$1012 to 12.590$1012 and then to
11.020$1012 when crossow velocity increased from 2.2 to 2.9 and
Fig. 5. Evolution of the resistances with time for sample S3 ultraltered at 3 bar and
then to 3.7 m s1, respectively. A few exceptions to this trend can be
2.2 m s1. Rt: total hydraulic resistance (m1); Rcf: cake layer resistance (m1); Ra:
adsorption and concentration polarization resistance (m1); Rm: membrane resistance
explained as a result of the different turbidity of the samples. For
(m1). the same operating conditions Ra increased when feed turbidity
increased, as it can be observed for samples S2 and S4 at 2 bar and
2.9 m s1 and for samples S1 and S3 at 3 bar and 3.7 m s1. On the
UF process. Several authors (Ho and Zydney, 2000; de la Casa et al., other hand, regarding the cake formation parameter, as crossow
2008; Mondal and De, 2010) reported that the evolution of velocity increased Rcf decreased, for all the transmembrane pres-
permeate ux with time can be divided in two stages: the rst one sures tested. This is explained by the reduction of the fouling layer
is caused by pore blocking and adsorption phenomena during the formed on the membrane surface during the UF process because of
C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23 21

the increase in the shear stress. Also, for a xed crossow velocity, 5.0
Rcf increased when transmembrane pressures increased. However,

Permeate flux 10 5 (m3m-2s-1)


Rcf for sample S1 ultraltered at 1 bar and 2.9 m s1 was greater
than that for samples S2 and S4 at 2 bar and 2.9 m s1. This is due to 4.0
Jp Resistance in Series model
the effect of turbidity, as sample S1 was the one with higher
turbidity. Also, for the same operating conditions Rcf increased 3.0
Jp Mondal resistance combined model
when turbidity increased, as it can be observed for sample S2 and
Jp (Sample S3; 3 bar; 2.2 m/s)
sample S4 at 2 bar and 2.9 m s1. For high TMP and low CFV Rcf had
similar values even for samples with very different turbidity. 2.0

4.2. Validation of resistance-in-series model 1.0

In order to validate the resistance-in-series model, the com-


bined resistance model proposed by Mondal and De was used, as it 0.0
was explained in Section 2.4 (Mondal and De, 2010). For this pur- 0.0 50.0 100.0 150.0 200.0
pose, the resistance of the cake layer previously determined by the Time (min)
resistance-in-series model was taken into account. Table 5 shows
Fig. 6. Comparison between the experimental permeate ux observed for sample S3 at
the tting accuracy results for the combined resistance model 3 bar and 2.2 m s1, the permeate ux predicted by the resistance-in-series model and
proposed by Mondal and De. Permeate ux predicted by the model the permeate ux predicted by the combined resistance model proposed by Mondal
proposed by Mondal and De was compared to the experimental and De (Mondal and De, 2010).
data and also to permeate ux predicted by the resistance-in-series
model. The tting accuracy results were expressed in terms of the
regression coefcient (R2) and standard deviation (SD). As it can be were tted to the experimental data of permeate ux decline with
observed, when model predictions are compared to the experi- time obtained with a 5 kDa PES UF membrane when different
mental data the accuracy was higher than 95% for all the runs samples of table olive production wastewaters were used as feed
except one (sample S1 ultraltered at 1 bar and 2.9 m s1). The solutions. Among all the models taken into account, the resistance-
standard deviation for the rest of the tests was between 0.040 and in-series model that considered pore blocking and absorption, and
0.022. The accuracy of the prediction considering the permeate ux cake formation fouling mechanisms was the most accurate in terms
data obtained with the resistance in series method was higher than of high R2 and low SD for all the transmembrane pressures, cross-
95% in all test with a standard deviation between 0.033 and 0.006. ow velocities and different feed solutions tested. Therefore, pore
This indicates that both models predicted similar values of blocking and absorption, and cake formation were the predominant
permeate ux and also the value Rcf is equivalent in both ones. Fig. 6 fouling mechanisms that explained the decline of permeate ux
compares permeate ux decline with time predicted by the with time.
resistance-in-series model to that predicted by the combined The inuence of transmembrane pressure and crossow veloc-
resistance method proposed by Mondal and De and to the experi- ity on the values of the resistance-in-series model parameters was
mental data for sample S3 ultraltered at 3 bar and 2.2 m s1. The determined. As transmembrane pressure increased the adsorption
gure shows very similar predictions for both models with a sharp and concentration polarization resistance (Ra) and the cake layer
initial reduction of permeate ux. Nevertheless, permeate ux resistance (Rcf) increased. As crossow velocity increased, Ra
predicted by the resistance-in-series model is closer to the exper- slightly decreased or keeps constant and Rcf decreased. As the
imental observations. turbidity of the feed sample increased Ra and Rcf increased. In the
steady state Rcf was the predominant fouling mechanism for the
operating conditions considered. The predominant fouling mech-
5. Conclusions anism in the rst minutes of the UF tests was Ra.
The combined resistance model proposed by Mondal and De,
Three different mathematical models (Hermia's models adapted 2010 was evaluated as well and the predictions were similar to
to crossow UF, a combined model and a resistance-in-series one)

Table 5
Fitting accuracy for the model developed by Mondal and De (Mondal and De, 2010) when the resistance-in-series parameters were considered: values of R2 and SD.

TMPa (bar) CFVb (m s1) Sample Feed turbidity (NTU) Mondal & De vs Mondal & De vs
experimental data resistance-in-series

R2 SDc R2 SDc

1 2.2 S2 314.3 0.966 0.034 0.997 0.006


2.9 S1 367.5 0.912 0.050 0.957 0.033
2 2.2 S2 314.3 0.976 0.029 0.991 0.015
2.9 S2 314.3 0.978 0.024 0.993 0.010
2.9 S4 183.4 0.976 0.022 0.993 0.009
3.7 S3 185.9 0.965 0.031 0.985 0.020
3.7 S4 183.4 0.963 0.023 0.992 0.010
3 2.2 S3 185.9 0.979 0.040 0.987 0.030
2.2 S1 367.5 0.974 0.036 0.993 0.016
2.9 S3 185.9 0.981 0.026 0.993 0.013
3.7 S3 185.9 0.975 0.025 0.989 0.015
a
TMP: transmembrane pressure.
b
CFV: crossow velocity.
c
SD: standard deviation.
22 C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23

that of the resistance-in-series model. It was observed that the cake dx.doi.org/10.1016/j.cej.2012.03.079.
Benitez, J.F., Acero, J.L., Leal, A.I., 2003. Purication of storage brines from the
layer resistance was equivalent for both models.
preservation of table olives. J. Hazard. Mater. 96, 155e169. http://dx.doi.org/
10.1016/S0304-3894(02)00183-8.
Acknowledgements Bodalo, A., Gomez, J.L., Go mez, M., Leo n, G., Hidalgo, A.M., Ruz, M.A., 2008. Phenol
removal from water by hybrid processes: study of the membrane process step.
Desalination 223, 323e329. http://dx.doi.org/10.1016/j.desal.2007.01.219.
The authors of this work wish to gratefully acknowledge the Bolton, G., LaCasse, D., Kuriyel, R., 2006. Combined models of membrane fouling:
nancial support of CDTI (Centre for Industrial Technological development and application to microltration and ultraltration of biological
uids. J. Membr. Sci. 277, 75e84. http://dx.doi.org/10.1016/
Development) depending on the Spanish Ministry of Science and
j.memsci.2004.12.053.
Innovation (INNPRONTA program, ITP-20111020). Carrere, H., Blaszkow, F., Roux de Balmann, H., 2001. Modelling the clarication of
lactic acid fermentation broths by cross-ow microltration. J. Membr. Sci. 186,
219e230. http://dx.doi.org/10.1016/S0376-7388(00)00677-3.
Nomenclature Carrere, H., Blaszkow, F., Roux de Balmann, H., 2002. Modelling the microltration
of lactic acid fermentation broths and comparison of operating modes. Desa-
lination 145, 201e206. http://dx.doi.org/10.1016/S0011-9164(02)00409-5.
Chen, H., Kim, A.S., 2006. Prediction of permeate ux decline in crossow mem-
List of symbols brane ltration of colloidal suspension: a radial basis function neural network
b fouling rate due to adsorption (s1) approach. Desalination 192, 415e428. http://dx.doi.org/10.1016/
K Hermia's model constant (units depending on n) j.desal.2005.07.045.
Cheryan, M., Alvarez, J.R., 1995. In: Noble, R.D. (Ed.), Food and Beverage Industry
Kc complete blocking model constant (s1) Applications, Membrane Separation Technology, Principles and Applications.
Kcf cake layer formation model constant (s m2) University of Colorado, Boulder, CO, USA, ISBN 978-0-444-81633-7 and S.A.
J permeate ux (m3 m2 s1) Stern, Syracuse University, Syracuse, NY, USA 341.
Jmodel permeate ux predicted by each model (m3 m2 s1) Choi, S.-W., Yoon, J.-Y., Haam, S., Jung, J.-K., Kim, J.-H., Kim, W.-S., 2000. Modeling of
the permeate ux during microltration of BSA-adsorbed microspheres in a
Jss steady-state permeate ux (m3 m2 s1) stirred cell. J. Colloid Interface Sci. 228, 270e278. http://dx.doi.org/10.1006/
n Hermia's model parameter (dimensionless) jcis.2000.6940.
Corbato n-B 
aguena, M.J., Vincent-Vela, M.C., Alvarez-Blanco, S., Lora-Garca, J., 2013.
P transmembrane pressure (bar)
Analysis of two ultraltration fouling models and estimation of model pa-
R total hydraulic resistance (m1) rameters as a function of operational conditions. Transp. Porous Media 99,
Ra resistance due to adsorption on membrane surface and 391e411. http://dx.doi.org/10.1007/s11242-013-0192-4.
n-B 
inside its pores and concentration polarization (m1) Corbato aguena, M.J., Alvarez-Blanco, S., Vincent-Vela, M.C., 2015a. Fouling
mechanisms of ultraltration membranes fouled with whey model solutions.
R'a steady-state adsorption and concentration polarization Desalination 360, 87e96. http://dx.doi.org/10.1016/j.desal.2015.01.019.
resistance (m1) Corbato n-B aguena, M.J., Gugliuzza, A., Cassano, A., Mazzei, R., Giorno, L., 2015b.
Rcf cake layer resistance (m1) Destabilization and removal of immobilized enzymes adsorbed onto poly-
ethersulfone ultraltration membranes by salt solutions. J. Membr. Sci. 486,
Rm new membrane resistance (m1) 207e214. http://dx.doi.org/10.1016/j.memsci.2015.03.061.
t ltration time (s) de Barros, S.T.D., Andrade, C.M.G., Mendes, E.S., Peres, L., 2003. Study of fouling
mechanism in pineapple juice clarication by ultraltration. J. Membr. Sci. 215,
213e224. http://dx.doi.org/10.1016/S0376-7388(02)00615-4.
Greek letters de la Casa, E.J., Guadix, A., Iba n~ ez, R., Camacho, F., Guadix, E.M., 2008. A combined
a fraction of membrane pores completely blocked fouling model to describe the inuence of the electrostatic environment on the
(dimensionless) cross-ow microltration of BSA. J. Membr. Sci. 318, 247e254. http://dx.doi.org/
10.1016/j.memsci.2008.02.047.
m feed solution viscosity (kg m1 s1) El-Abbassi, A., Kiai, H., Raiti, J., Hadi, A., 2014. Application of ultraltration for olive
processing wastewaters treatment. J. Clean. Prod. 65, 432e438. http://
Abbreviations dx.doi.org/10.1016/j.jclepro.2013.08.016.
Field, R.W., Wu, D., Howell, J.A., Gupta, B.B., 1995. Critical ux concept for micro-
CFV crossow velocity (m s1) ltration fouling. J. Membr. Sci. 100, 259e272. http://dx.doi.org/10.1016/0376-
COD chemical oxygen demand 7388(94)00265-Z.
MWCO molecular weight cut off Galanakis, C.M., Tornberg, E., Gekas, V., 2010. Clarication of high-added value
products from olive mill wastewater. J. Food Eng. 99, 190e197. http://dx.doi.org/
NF nanoltration 10.1016/j.jfoodeng.2010.02.018.
NTU nephelometric turbidity unit Garrido Fern andez, A., Ferna ndez Dez, M.J., Adams, R.M., 1997. Table Olives. Pro-
OMW olive mill wastewater duction and Processing. Chapman & Hall, UK London, ISBN 978-0412718106.
Hermia, J., 1982. Constant pressure blocking ltration laws e application to pow-
PES polyethersulfone erlaw non-newtonian uids. Inst. Chem. Eng. 60, 183e187.
SD standard deviation Ho, C.-C., Zydney, A.L., 2000. A combined pore blockage and cake ltration model
TMP transmembrane pressure (bar) for protein fouling during microltration. J. Colloid Interface Sci. 232, 389e399.
http://dx.doi.org/10.1006/jcis.2000.7231.
TOPP table olive packing plant
International Olive Oil Council, 2014. Compiled Data Taken from the Total Produced
TSS total suspended solids Worldwide of Total Table Olives by Country.
UF ultraltration Karasu, K., Yoshikawa, S., Ookawara, S., Ogawa, K., Kentish, S.E., Steves, G.W., 2010.
A combined model for the prediction of the permeation ux during the cross-
ow ultraltration of a whey suspension. J. Membr. Sci. 361, 71e77. http://
References dx.doi.org/10.1016/j.memsci.2010.06.008.
Lin, C.-J., Rao, P., Shirazi, S., 2004. Effect of operating parameters on permeate ux
Alburquerque, J.A., Gonzalvez, J., Garca, D., Cegarra, J., 2004. Agrochemical char- decline caused by cake formation e a model study. Desalination 171, 95e105.
acterisation of alperujo, a solid by-product of the two-phase centrifugation http://dx.doi.org/10.1016/j.desal.2004.03.023.
method for olive oil extraction. Bioresour. Technol. 91, 195e200. http:// Lin, S.-H., Hung, C.-L., Juang, R.-S., 2008. Applicability of the exponential time
dx.doi.org/10.1016/S0960-8524(03)00177-9. dependence of ux decline during dead-end ultraltration of binary protein
Astaree, R.S., Mohammadi, T., Kasiri, N., 2015. Analysis of BSA, dextran and humic solutions. Chem. Eng. J. 145, 211e217. http://dx.doi.org/10.1016/
acid fouling during microltration, experimental and modeling. Food Bioprod. j.cej.2008.04.003.
Process 94, 331e341. http://dx.doi.org/10.1016/j.fbp.2014.04.003. Liu, H., Tang, Z., Cui, C., Sun, C., Zhu, H., Li, B., Guo, L., 2014. Fouling mechanisms of
Astudillo-Castro, C., 2015. Limiting ux and critical transmembrane pressure the extract of traditional Chinese medicine in ultraltration. Desalination 354,
determination using an exponential model: the effect of concentration factor, 87e96. http://dx.doi.org/10.1016/j.desal.2014.09.016.
temperature, and cross-ow velocity during casein micelle concentration by Mah, S.-K., Chuah, C.-K., Cathie Lee, W.P., Cahi, S.-P., 2012. Ultraltration of palm oil-
microltration. Ind. Eng. Chem. Res. 54, 414e425. http://dx.doi.org/10.1021/ oleic acid-glycerin solutions: fouling mechanism identication, fouling mech-
ie5033292. anism analysis and membrane characterizations. Sep. Purif. Technol. 98,
Barredo-Damas, S., Alcaina-Miranda, M.I., Iborra-Clar, M.I., Mendoza-Roca, J.A., 419e431. http://dx.doi.org/10.1016/j.seppur.2012.07.020.
2012. Application of tubular ceramic ultraltration membranes for the treat- Miller, D.J., Kasemset, S., Paul, D.R., Freeman, B.D., 2014. Comparison of membrane
ment of integrated textile wastewaters. Chem. Eng. J. 192, 211e218. http:// fouling at constant ux and constant transmembrane pressure conditions.
C. Carbonell-Alcaina et al. / Journal of Food Engineering 187 (2016) 14e23 23

J. Membr. Sci. 454, 505e515. http://dx.doi.org/10.1016/j.memsci.2013.12.027. Desalination 274, 262e271. http://dx.doi.org/10.1016/j.desal.2011.02.015.


Mondal, S., Cassano, A., Tasselli, F., De, S., 2011. A generalized model for clarication Ruby Figueroa, R.A., Cassano, A., Drioli, E., 2011. Ultraltration of orange press li-
of fruit juice during ultraltration under total recycle and batch mode. quor: optimization for permeate ux and fouling index by response surface
J. Membr. Sci. 366, 295e303. http://dx.doi.org/10.1016/j.memsci.2010.10.015. methodology. Sep. Purif. Technol. 80, 1e10. http://dx.doi.org/10.1016/
Mondal, S., De, S., 2010. A fouling model for steady state crossow membrane j.seppur.2011.03.030.
ltration considering sequential intermediate pore blocking and cake forma- Tien, C., Ramarao, B.V., Yasarla, R., 2014. A blocking model of membrane ltration.
tion. Sep. Purif. Technol. 75, 222e228. http://dx.doi.org/10.1016/ Chem. Eng. Sci. 111, 421e431. http://dx.doi.org/10.1016/j.ces.2014.01.022.
j.seppur.2010.07.016. Turano, E., Curcio, S., De Paola, M.G., Calabro , V., Iorio, G., 2002. An integrated
Mondal, S., De, S., 2009. Generalized criteria for identication of fouling mechanism centrifugation-ultraltration system in the treatment of olive mil wastewater.
under steady state membrane ltration. J. Membr. Sci. 344, 6e13. http:// J. Membr. Sci. 209, 519e531. http://dx.doi.org/10.1016/S0376-7388(02)00369-1.
dx.doi.org/10.1016/j.memsci.2009.08.015. 
Vincent Vela, M.C., Alvarez Blanco, S., Lora Garca, J., Bergatin~ os Rodrguez, E., 2009.
Muthukumaran, S., Kentish, S.E., Ashokkumar, M., Stevens, G.W., 2005. Mechanisms Analysis of membrane pore blocking models adapted to crossow ultraltration
for the ultrasonic enhancement of dairy whey ultraltration. J. Membr. Sci. 258, in the ultraltration of PEG. Chem. Eng. J. 149, 232e241. http://dx.doi.org/
106e114. http://dx.doi.org/10.1016/j.memsci.2005.03.001. 10.1016/j.cej.2008.10.027.
Paraskeva, C.A., Papadakis, V.G., Tsarouchi, E., Kanellopoulou, D.G., Koutsoukos, P.G., 
Vincent-Vela, M.C., Cuartas-Uribe, B., Alvarez-Blanco, ~ os-
S., Lora-Garca, J., Bergantin
2007. Membrane processing for olive mill wastewater fractionation. Desalina- Rodrguez, E., 2010. Analysis of ultraltration processes with dilatant macro-
tion 213, 218e229. http://dx.doi.org/10.1016/j.desal.2006.04.087. molecular solutions by means of dimensionless numbers and hydrodynamic
nchez-Monedero, M.A., Bernal, M.P., 1999. Char-
Paredes, C., Cegarra, J., Roig, A., Sa parameters. Sep. Purif. Technol. 75, 332e339. http://dx.doi.org/10.1016/
acterization of olive mill wastewater (alpechin) and its sludge for agricultural j.seppur.2010.09.001.
purposes. Bioresour. Technol. 67, 111e115. http://dx.doi.org/10.1016/S0960- Wang, L., Song, L., 1999. Flux decline in crossow microltration and ultraltration:
8524(98)00106-0. experimental verication of fouling dynamics. J. Membr. Sci. 160, 41e50. http://
Peng, H., Tremblay, A.Y., 2008. Membrane regeneration and ltration modeling in dx.doi.org/10.1016/S0376-7388(99)00075-7.
treating oily wastewaters. J. Membr. Sci. 324, 59e66. http://dx.doi.org/10.1016/ Yee, K.W.K., Wiley, D.E., Bao, J., 2009. A unied model of the time dependence of
j.memsci.2008.06.062. ux decline for the long-term ultraltration of whey. J. Membr. Sci. 332, 69e80.
Rezaei, H., Ashtiani, F.Z., Fouladitajar, A., 2011. Effects of operating parameters on http://dx.doi.org/10.1016/j.memsci.2009.01.041.
fouling mechanism and membrane ux in cross-ow microltration of whey.

Das könnte Ihnen auch gefallen