You are on page 1of 9

Innovative Food Science and Emerging Technologies 12 (2011) 352–360

Contents lists available at ScienceDirect

Innovative Food Science and Emerging Technologies


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / i f s e t

Athermal concentration by osmotic evaporation of roselle extract, apple and grape


juices and impact on quality
Mady Cissé a, Fabrice Vaillant b,c, Sonia Bouquet c, Dominique Pallet c, Florence Lutin d,
Max Reynes c, Manuel Dornier c,⁎
a
Ecole Supérieure Polytechnique, Université Cheikh Anta Diop, B.P. 5085, Dakar Fann, Sénégal
b
Centro de Investigación de Tecnología de Alimentos, Universidad de Costa Rica, Codigo Postal 2060, San José, Costa Rica
c
CIRAD, Montpellier SupAgro, UMR Qualisud, 73 rue J.F. Breton, TA B-95/16, F-34398 Montpellier cedex 5, France
d
EURODIA Industries S.A., 7 rue du Jura, BP. 30527 Wissous, F-94633 Rungis cedex, France

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

Article history: Osmotic evaporation (or osmotic distillation) was carried out on roselle extract, apple and grape juices. The
Received 21 January 2011 industrial pilot plant used had a hydrophobic, polypropylene, hollow-fiber membrane with an area of 10.2 m2
Accepted 21 February 2011 and an average pore diameter of 0.2 μm. It was suitable for concentrating vegetable extracts and fruit juices,
and controlled various parameters such as temperature, flow velocity, and brine concentration. The final total
Keywords:
soluble solids (TSS) contents achieved were 660, 570, and 610 g kg−1 for grape juice, apple juice, and roselle
Osmotic evaporation
Concentration
extract, respectively. Temperature and concentration of solutions significantly influenced evaporation flux,
Membrane contactor which, for roselle extract, was 1.5 kg h−1 m−2 at 610 g TSS kg−1 and 45 °C. The physico-chemical,
Hibiscus sabdariffa biochemical, and aromatic qualities of concentrates obtained by osmotic evaporation were much higher
Grape juice than those of thermal concentrates, and close to those of the initial products.
Apple juice Industrial relevance: Membrane processes are increasingly used to concentrate thermo-sensitive fruit juices
Product quality and plant extracts. Their capacity to operate at moderate temperatures and pressures means that their energy
consumption is low, while they produce good quality concentrates. Nonetheless, the main disadvantage of
baromembrane processes is their inability to reach the concentration levels standard for products of thermal
evaporation because of limitations resulting from high osmotic pressure. Actually, reverse osmosis
membranes and equipment limit the final concentration of fruit juices to about 25–35°Brix. Osmotic
evaporation has attracted considerable interest, as it can concentrate juices to as much as 65°Brix. This
process, when applied to various juices, better preserves the quality of raw materials. However, because of the
geometrical limitations of commercially available membranes and modules, juices must first be clarified. To
our knowledge, only a few studies on osmotic evaporation have so far been conducted at a semi-industrial
scale and never with roselle extracts.
© 2011 Elsevier Ltd. All rights reserved.

1. Introduction Research on technological alternatives to thermal processing has


become an important issue. Membrane processes such as ultrafiltra-
The preservation, during stabilization, of nutritional, sensorial, tion, nanofiltration, reverse osmosis, and osmotic evaporation are
antioxidant, and pharmaceutical properties of fruit juices and plant today increasingly used to concentrate fruit juices and plant extracts.
extracts has become a major challenge in food science. Classical Their capacity to operate at moderate temperatures and pressures
processes, such as thermal pasteurization and concentration by means that their energy consumption is low, while they produce good
vacuum evaporation, significantly change the quality of fresh fruit quality concentrates (Banvolgyi, Horvath, Stefanovits-Banyai,
juices and plant extracts. Temperatures higher than 50 °C degrade Bekassy-Molnar, & Vatai, 2009; Cassano, Jiao, & Drioli, 2004; Ding,
sensorial properties and nutritional compounds such as vitamins, and Liu, Yu, Ma, & Yang, 2008; Hongvaleerat, Cabral, Dornier, Reynes, &
induce a loss of aroma compounds, leading to a partial loss of the fresh Ningsanond, 2008). Nonetheless, the main disadvantage of baromem-
juice flavor (Cisse, Vaillant, Perez, Dornier, & Reynes, 2005; Shaw brane processes is their inability to reach the concentration levels
et al., 2001; Vaillant et al., 2001). standard for products of thermal evaporation because of limitations
resulting from high osmotic pressure. Actually, reverse osmosis
membranes and equipment limit the final concentration of fruit juices
to about 25–35°Brix (Banvolgyi et al., 2009; Rodrigues et al., 2004).
⁎ Corresponding author. Tel.: +33 467617186; fax: +33 467614449. Osmotic evaporation (OE), also called osmotic distillation, has
E-mail address: dornier@cirad.fr (M. Dornier). attracted considerable interest (Girard & Fukumoto, 2000; Hogan,

1466-8564/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ifset.2011.02.009
M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360 353

Canning, Peterson, Johnson, & Michaels, 1998; Jiao, Cassano, & Drioli, a polyester bag-filter system with a micron rating of 5 μm (GAF,
2004), as it can concentrate juices to as much as 650 g total soluble Belgium). The extract was then stored at 4 °C until used. Industrial
solids (TSS) per kilogram (Cisse et al., 2005; Rodrigues et al., 2004; pasteurized apple and grape juices were purchased. Both juices were
Vaillant et al., 2001). This process, when applied to various juices, completely clear and were probably clarified. In accordance with the
better preserves the quality of raw materials (Cisse et al., 2005; European legislation, these juices only contained the aroma com-
Hongvaleerat et al., 2008; Vaillant et al., 2005). However, because of pounds of the fruits.
the geometrical limitations of commercially available membranes and
modules, juices must first be clarified. To our knowledge, only a few 2.2. Concentration
studies on osmotic evaporation have so far been conducted at a semi-
industrial scale (Cisse et al., 2005; Vaillant et al., 2001; Vaillant et al., An industrial pilot plant, developed by EURODIA Industrie
2005) and never with roselle extracts. (Wissous, France) and CIRAD, was used (Fig. 1). The plant featured
Roselle (Hibiscus sabdariffa L.) is a herbaceous plant cultivated a hydrophobic, polypropylene, hollow-fiber membrane with a total
largely in the tropics and subtropics of both hemispheres (Cisse, area of 10.2 m2 and an average pore diameter of 0.2 μm. The supply
Dornier, Sakho, Ndiaye, Reynes & Sock, 2009). Belonging to the tank for the juice had a 50 L capacity, while the brine circuit had a total
Malvaceae family, H. sabdariffa is known under different names, such volume of 70 L. The juice or extract to be concentrated circulated
as bissap in Senegal, karkade in North Africa, roselle or sorrel in English, inside the hollow fibers.
and flor de Jamaica in Central and South America. Aqueous extract from The concentrate loop was continuously fed with roselle extract or
calyces of H. sabdariffa presents several interesting features. It is used fruit juice, and the concentrate extracted continuously, once the
worldwide to produce drinks and is a source of natural food coloring desired level of total soluble solids was reached. Calcium chloride
because of its high anthocyanin content (Cisse, Dornier, et al., 2009; solution, that is, the brine solution, circulated concurrently on the
Juliani et al., 2009). Hibiscus extracts are reported as having medical other side of the membrane. Electrical conductivity of the juice was
properties such as decreasing serum cholesterol in humans and continuously monitored during concentration to ensure membrane
animals (Hirunpanich et al., 2006; Lin et al., 2007), protecting the liver integrity and hydrophobicity, and detect any possible salt leakage
against oxidation stress (Liu et al., 2006), having antihypertensive and through the membrane. During the trials, CaCl2 crystals were added at
cardioprotective effects (Odigie, Ettarh, & Adigun, 2003), and 5.5 to 6.0 mol L−1 to maintain the brine solution at near saturation.
attenuating nephropathy in diabetes (Lee et al., 2009). Most According to Cisse et al. (2005) and Vaillant et al. (2001) this
nutritional and functional properties of Roselle extracts can be configuration better preserves the concentrate's aroma compounds.
attributed to its very high content of anthocyanin acting probably in Brine temperature was maintained between 37 and 40 °C by cooling.
synergy with other compounds. The main problem, during thermal The temperature of solution to be concentrated was maintained at
processing of Roselle, is that high temperatures drastically degrade 35 ± 2 °C. Pressure and temperature values at the membrane's inlet
anthocyanins and also reduce their stability during storage (Cisse, and outlet were recorded, pressure varying by ± 2% and temperature
Vaillant, et al., 2009). by ± 1 °C. The average evaporation flux (Jw) was measured on the
This paper evaluates the potential of using osmotic evaporation to brine loop.
concentrate roselle extract, apple and grape juices. The process is The brine tank had an opening in the upper layer and water,
evaluated in terms of performance and impact on product quality and evaporated from the product being concentrated, condensed in the
compared with vacuum evaporation. brine, resulting in increased brine volume. Brine overflow was then
collected in a container placed on a balance (PRECIA MOLEN SA, Privas,
2. Materials and methods France) with a maximum capacity of 30 ± 0.01 kg. Mass differences,
reported regularly every 5 min between two successive measure-
2.1. Raw materials ments, gave the evaporation flux. At the end of the trials, the diluted
brine was recovered and concentrated by heating at atmosphere
To prepare roselle extract, dried calyces of Senegalese-grown Thai pressure until salts crystallized. The crystals were kept for reuse in
variety were mixed with cooled demineralized water at a mass ratio further trials.
of calyces to water at 1:5. After soaking for 3 h, the extract was To compare osmotic evaporation with a reference thermal process,
filtered, first through a stainless steel sieve (1 mm) and then through roselle extract, grape and apple juices were also concentrated by

Fig. 1. Schematic of the pilot plant of osmotic evaporation used for concentration of roselle extract, apple and grape juices.
354 M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360

evaporation at 80 °C under a 50-kPa vacuum, using a vacuum carrier gas, with a constant column flow rate of 1 mL min−1. The
evaporator with a 10-L capacity and equipped with a stirrer (H.P. transfer line temperature was kept at 280 °C.
Auriol, Marmande, France). This apparatus allows to simulate the The mass spectrometer used to identify headspace compounds
operating conditions in the first effect of a classical industrial vacuum- operated in electron ionization mode (internal ionization at 70 eV)
evaporator without aroma recovery system. with a scan range from m/z 40 to 400. Compound identifications were
performed, using either NIST (2002 version) or the Wiley 275 spectral
library. The spectra data from authentic compounds plus the linear
2.3. Analyses
retention index were compared with those found in the literature.
Linear retention indices (type Kovats) were calculated after analysis,
All extracts and juices were analyzed for pH, density, and dry matter,
under the same chromatographic conditions, of the C8-C20 n-alkane
using standard methods (AOAC. 1990). Total soluble solid (TSS)
series (Supelco, Inc., Bellefonte, PA, USA) according to (Drosos, Viola-
contents were measured, using a digital PAL-3 refractometer (Atago
Rhenals, & Vivas-Reyes, 2010). Another internal standard for
Co., Japan) with automatic temperature compensation. Total titratable
quantification was 4-methyl-1-pentanol from Sigma (Steinheim,
acidity was determined with a TitroLine easy tritator (SCHOTT
Germany).
Instruments, Mainz, Germany), using 0.1 N NaOH. Vitamin C (ascorbic
The quantitative assay was performed by comparing the area of
acid and dehydroascorbic acid) was assessed by a high-performance
analyte peaks with that of the internal standard. Losses of aroma
liquid chromatography, as according to Dhuique-Mayer et al. (2007)
compounds and aromatic distance (AD), expressed in arbitrary units,
and using a 1100 HPLC System (Agilent Technologies, Massy, France).
were evaluated, using Eqs. 1 and 2.
Sucrose, fructose, and glucose were determined, using Dionex DX-
600 high-performance liquid chromatography, fitted with a Carbo-  
Pac® MA1 column. The eluant was sodium hydroxide solution at a Af −Ai
% losses =  100 ð1Þ
concentration of 0.6 mol L−1 and flow rate of 0.4 mL min−1. After Ai
separation, sugars were detected, using a Dionex ED50 pulsed
amperometric detector. The total sugar content corresponded to the rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi 2
sum of the contents of the different carbohydrate compounds AD = ∑ Ai −Af ð2Þ
detected in the kernel. Before analysis, fruit juices and roselle extracts
were diluted and filtered through 0.45-μm filters.
Where, Ai and Af are, respectively, the area of aroma compounds
Total anthocyanin content was assessed with the pH differential
before and after osmotic evaporation (or vacuum evaporation), and
method (Lee, Durst, & Wrolstad, 2005). All absorbance readings were
AD is the distance from the initial product in an aromatic space of n
made against distilled water, which acted as control. Spectrophoto-
dimensions.
metric measurements were carried out, using UV-1605 spectro-
photometers (Shimadzu, Kyoto, Japan). Concentrations were
2.5. Statistical analyses
expressed as delphinidin 3-xylosylglucoside equivalents for roselle
(MW = 577 g mol−1). The molar extinction coefficient at pH 1 and
All statistical analyses were performed using XLSTAT software
510 nm, used for calculation, was 26,000 L mol−1 cm−1 (Cisse,
(Microsoft Corp.).
Vaillant, et al., 2009). All the reagents used were of analytical grade
and were purchased from Sigma (l'Isle d'Abeau, France). The total
3. Results and discussion
phenolic content was determined with Folin–Ciocalteu reagent,
according to the method optimized by George, Brat, Alter, & Amiot
3.1. Evaporation fluxes
(2005).
Viscosity was measured at 35 °C, using a HAAKE VT550 viscometer
Table 1 lists the average operating conditions during the osmotic
(Thermo, Berlin, Germany), fitted with a coaxial cylinder sensor
evaporation, on a semi-industrial scale, of water, sucrose solutions at
(sensor system NV). Water activity was measured with an AquaLab
100 g TSS kg−1, roselle extracts, apple juice, and grape extract. Also
4TE water activity meter (Decagon Devices, Inc., Pullman, WA, USA) at
shown are the resulting final concentrations achieved (TSSf) and the
35 °C. Alcohol insoluble solids (AIS) such as pectins were determined
average evaporation fluxes. The equipment developed enabled
by dispersing a previously weighed homogenous aliquot of retentate
accurate control of all operating conditions and, consequently, when
into boiling ethanol (80 vol.%) (Vaillant et al., 1999).
replicates of the trial were done, a very high reproducibility of final
Oxygen radical absorbance capacity (ORAC) assay was performed in
results was achieved, even for different products.
accordance with Ou, Hampsch-Woodill, & Prior (2001), using a
Because of pilot configuration, brine temperature was maintained
microplate spectrofluorimeter with 96-well plates made in black
at least 2 °C above the temperature of the solution being concentrated.
polypropylene (Gancel, Alter, Dhuique-Mayer, Ruales, & Vaillant,
For all trials, the average evaporation flux ranged between 1.13 and
2008). The ORAC value was expressed as Trolox equivalents (μmol g−1
1.48 kg h−1 m−2 and appeared to depend only on the juice concen-
FW).
tration and temperature of the solution to be concentrated.
The first set of experiments was carried at 35 ± 2 °C with demin-
2.4. Headspace solid-phase microextraction (HS-SPME) of free volatiles eralized water and sucrose solution at 100 g TSS kg−1 to assess the
evaporation performance of the pilot plant. Fig. 2 presents the
Juice and extract samples were analyzed with an Agilent 6890N evaporation fluxes obtained for both water (fluctuating between
gas chromatograph, which was equipped with a mass spectrometer 1.40 and 1.50 kg h−1 m−2) and sugar solution (between 1.11 and
Agilent MSD 5973N as detector. Free volatiles were adsorbed on SPME 1.41 kg h−1 m−2), as measured against time. These evaporation
fiber (polydimethylsiloxane-divinylbenzene, i.e., PDMS-DVB; 65 μm) fluxes are, on average, seven times smaller than those obtained by
and desorbed in splitless mode for 3 min in the GC injector at 250 °C. Courel, Dornier, Herry, Rios, and Reynes (2000), who studied osmotic
Separation was achieved in a J&W DB-WAX 122-7032 column (30 m, evaporation in the laboratory, using flat, thin, porous membranes of
0.25 mm inner diameter, 0.25-μm film thickness, fused-silica liner; polytetrafluoroethylene. However, they were 50 to 100% higher than
Agilent Technologies, Palo Alto, CA, USA). The GC oven temperature the results of other trials, where the same module and membranes
program consisted of 40 °C for 3 min, 40 to 170 °C at 3 °C min−1, 170 were used (Ali, Dornier, Duquenoy, & Reynes 2003; Cisse et al.,
to 240 °C for 10 °C min−1, and 240 °C for 10 min. Helium was used as a 2005).
M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360 355

Table 1
Average operating conditions (T temperatures; ΔT = Tbrine − Tjuice; C concentration; and TSS total soluble solids) and water flux (Jw) of the different trials of osmotic evaporation
performed on the industrial pilot plant. () Standard deviation of n trials.

Product Tjuice Tbrine ΔT Cbrine TSSinitial TSSfinal Jw


(°C) (°C) (°C) (mol L−1) (g.kg−1) (g.kg−1) (kg h−1 m−2)

Water 35.1 38.5 3.5 5.9 0.0 0.0 1.45


(n = 4) (0.7) (0.7) (0.4) (0.1) (0.01)
Sucrose solution 34.8 37.4 2.6 5.6 106.7 539 1.29
(n = 3) (0.1) (0.1) (0.1) (0.1) (2.9) (14.5) (0.02)
Sucrose solution 34.8 37.2 2.4 5.6 150 650 1.26
Roselle extract 37.4 40.2 2.8 5.6 90 548 1.15
(n = 2) (0.1) (0.1) (0.1) (0.1) (1.0) (1.0) (0.01)
Roselle extract 32.1 34.6 2.5 5.6 90 535 1.13
Roselle extract 34.9 38.2 3.3 5.6 90 557 1.16
Roselle extract 44.9 48.2 3.3 5.6 90 615 1.48
Apple juice 35.3 37.9 2.6 5.6 110 570 1.17
Grape juice 35.2 37.5 2.3 5.6 157 660 1.17

Average evaporation fluxes (Jw) were slightly higher for water almost constant at 5.3 ± 0.1 mol L−1 during the last part of the trial.
than for sucrose solution, respectively at 1.43 and 1.26 kg h−1 m−2 The significant decrease of evaporation flux corresponded to an
after 60 min of operation. These findings may be attributed to a exponential increase of the solution's viscosity (Fig. 4), which may
polarization effect. Indeed, for sucrose solution, TSS content increased raise resistance to mass transfer in the liquid phase and, consequently,
from 105 to 538 g TSS kg−1. to a polarization effect, which may then induce a lower driving force.
Using similar operating conditions, a roselle extract was concentrat- Similar results were obtained for sucrose solution, and pineapple,
ed from 90 to 543 g TSS kg−1 in 10.5 h. Although, the hold-up volume of citrus, carrot, kiwifruit, and camu-camu juices in the laboratory
the concentrate loop was minimized, this lengthy time results from the (Cassano, Figoli, Tagarelli, Sindona, & Drioli, 2006, Cassano et al., 2003;
high ratio between the hold-up volume and the average evaporation flux Courel, Dornier, Herry, et al., 2000; Hongvaleerat et al., 2008;
(about 1.2 kg h−1 m−2). The behavior of evaporation flux against time Rodrigues et al., 2004); and for passion fruit, orange, and melon
can be observed in Fig. 3a. When TSS increased, the evaporation flux juices on a semi-industrial scale (Cisse et al., 2005; Vaillant et al.,
decreased from 1.30 kg h−1 m−2 at low TSS to 1.13 kg h−1 m−2 when 2001; Vaillant et al., 2005). On the basis of these results, at low TSS, as
TSS reached 400 g kg−1 and to 0.78 kg h−1 m−2 when TSS reached
543 g kg−1. An inflexion point of evaporation flux was observed when
concentration level reached 400 g kg−1, corroborating previously
reported results (Courel, Dornier, Herry, et al., 2000).
During the first part of the trial, the decline in evaporation flux was
slight and could be attributed to a small decrease in brine concen-
tration from 5.7 mol L−1 at TSS of 90 g kg−1 to 5.3 mol L−1 when TSS
reached 400 g kg−1 (Fig. 4). This result shows the relatively important
influence that brine concentration has on evaporation flux. Similar
results were observed by other authors (Courel, Dornier, Herry, et al.,
2000; Vaillant et al., 2001). During the second part of trial (between
400 and 543 g kg−1) (Fig. 3a), evaporation flux declined more rapidly
until a minimum of 0.78 kg h−1 m−2 was reached when concentra-
tion reached 543 g kg−1, corresponding to a total decrease of flux by
40% when comparing with initial flux value.
This result shows the strong influence that concentration level has
on evaporation flux, considering that brine concentration remained

Fig. 3. Evolution of water flux and total soluble solids during the osmotic evaporation of
Fig. 2. Water flux and concentration in total soluble solids (TSS) during osmotic evap- the roselle extract (a) Tj = 34.9 °C and Tb = 38.2 °C; and (b) Tj = 44.9 °C and
oration of water and sugar solution. Tb = 48.2 °C.
356 M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360

Fig. 5. Evolution of water flux and total soluble solids during the osmotic evaporation of
Fig. 4. Evolution of brine concentration and extract viscosity during osmotic apple and grape juice (Tj = 34 ± 2 °C and Tb = 36 ± 2 °C °C).
evaporation of roselle extract at Tj = 34.9 °C and Tb = 38.2 °C.

(r2 = 0.928).
in the case of water, evaporation flux can be said to depend mainly on
brine concentration. At concentration values higher than Jw = 0:203:LnðReÞ + 0:272 ð3Þ
40 g TSS 100 g−1, evaporation rates depend predominantly on solu-
tion viscosity and consequently on concentration level and temper-
In our specific case, this model may be used to predict osmotic
ature. Because evaporation flux appears to be independent of time, we
evaporation flux for different solution by knowing only the Reynolds
can assume that no significant membrane fouling occurred in any of
number.
the trials.
Fig. 3b shows the results obtained during a trial similar to the ones
3.2. Quality of concentrates obtained
described above, except that the temperature of roselle extract and
brine solution was increased by almost 10 °C. Evaporation flux
Table 2 presents the main physico-chemical properties of the
increased significantly with respect to the previous trials from
initial roselle extract, and apple and grape juices; and their
1.42 kg h−1 m−2 at low TSS to 1.76 kg h−1 m−2, and from 0.78 to
concentrates as obtained by either osmotic evaporation (OE) or
1.24 kg h−1 m−2 when TSS reached 610 g kg−1. Indeed, several
vacuum evaporation (VE). For all products, pH did not change because
authors (Courel, Dornier, Herry, et al., 2000; Courel, Dornier, Rios, &
of the buffering capacity of the fruit juices and roselle extract.
Reynes, 2000; Hogan et al., 1998; Jiao et al., 2004) have shown that for
Comparisons expressed on a dry matter basis (g kg−1 DM) of solute
temperatures higher than ambient temperature, the performance of
content (sugar, pectins, anthocyanins, and polyphenols) show,
osmotic evaporation is significantly enhanced as viscosity drops and
overall, no significant differences between the initial product and
solute diffusivity increases, thus inducing a higher driving force.
OE concentrate for all juices and roselle extract.
In our case, increasing temperatures by 10 °C resulted in a global
However, significant differences were noted between the initial
increase of evaporation flux by almost 40%. This increase would have
products and the VE concentrates. For example, for apple juice, total
been greater if the temperature differential between solution and
acidity was 20% lower for the VE concentrate than for the raw juice.
brine had been positive (Tj N Tb). However, because of equipment
Glucose contents in the VE concentrate were 51 and 38% lower than in
limitations, a negative temperature differential was maintained, even
the initial products of grape and roselle, respectively. Similar results
though, according to previous studies, the evaporation flux may have
were observed during the thermal concentration of orange, melon,
been increased by a factor of two for a positive temperature
and passion fruit juices (Cisse et al., 2005; Vaillant et al., 2001; Vaillant
differential of 12 to 15 °C between the extract and brine (Courel,
et al., 2005). These decreases in sugar and organic acid content can be
Dornier, Herry, et al., 2000; Hongvaleerat et al., 2008).
explained by several reactions such as sucrose inversion resulting
Osmotic evaporation at 35 °C, using the same equipment, was
from lower acidity and temperature, acid degradation of hexoses, and
applied to commercial, clarified, and pasteurized juices of apple and
oxidation.
grape. Fig. 5 shows the evaporation flux obtained versus time, when
Losses of anthocyanins were slight but not significant for the OE
the grape and apple juices were concentrated, respectively, from 150
concentrates and only 6% for the VE concentrates. These values
to 660 g TSS kg−1 and 110 to 570 g TSS kg−1. A similar trend was
obtained for roselle extract, with the average evaporation flux ranging
between 1.10 and 1.22 kg h−1 m−2.
To determine the conditions under which the mass transfer
resistance of the boundary layer can be ignored, the evaporation
fluxes obtained during the different trials with all the solutions tested
were reported as a function of the Reynolds number. This number is
calculated by taking into account viscosity of the solutions, flow inside
module, and the geometry of the hollow fibers (Fig. 6). For a Reynolds
number higher than 100, evaporation flux does not increase any
further, in contrast to what happens for lower Reynolds numbers. This
result indicates that, according to our operating conditions and
configuration of membranes and module, the main resistance to mass
transfer is in the liquid phase when Reynolds numbers are less than
100 and primarily in the membrane when Reynolds numbers are high. Fig. 6. Evaporation flux versus Reynolds number in the hollow fibers of the concentrate
A simple model (Eq. 3) can fit these results, with good correlation loop.
M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360 357

Table 2
Comparison of the main characteristics of the initial product and the concentrates obtained by osmotic evaporation (OE) and vacuum evaporation (VE); for a product value with the
same letter is not significantly different.

Apple juice Grape juice Roselle extract

Initial OE VE Initial OE VE Initial OE VE


a a a
pH 3.3 3.4 3.5 3.0 3.0 3.4 2.3 2.4 2.4
Density at 35 °C 1099a 1266b 1263b 1058 1316 1303 1033 1283 1290
(kg m−3)
Total soluble solids 112a 570b 565c 159 668 682 91 549 560
(g kg−1)
Dry matter 111a 572b 566c 161 669 683 92 551 561
(g kg−1)
Titratable acidity
(mEq kg−1) 63 320 260 74 301 257 505 3000 2274
(mEq kg−1 DM) 568a 560b 459c 460a 450b 387c 5489a 5446a 4053b
Glucose
(g kg−1) 22 112 85 79 327 163 21 123 81
(g kg−1 DM) 201a 196a 151b 491a 489a 239b 232a 223a 144b
Fructose
(g kg−1) 68 334 230 83 330 255 16 78 55
(g kg−1 DM) 613a 583b 406a 516a 493b 373b 174a 142b 98c
Sucrose
(g kg−1) 19 96 54 Traces Traces Traces 46 248 4
(g kg−1 DM) 171a 168a 95b 500a 450b 7c
Pectin (g kg−1 DM) 0 0 0 0 0 0 473a 466a 464a
Viscosity at 35 °C 1.3a 18.7b 13.7c 1.6a 40.6b 39.9b 1.6a 44.0b 88.8c
(mPa s)
Aw at 35 °C 0.988a 0.864b 0.865b 0.982a 0.780b 0.812c 0.990a 0.902b 0.912c
Polyphenols
(mg kg−1) 368 1647 1617 1061 4322 3940 2644 15756 25320
(g kg−1DM) 3.3a 2.9ab 2.9b 6.6a 6.5a 5.8b 28.7a 28.6a 45.1b
Anthocyanins
(mg L−1) nd nd nd nd nd nd 240 1400 1380
(g kg−1 DM) nd nd nd nd nd nd 2.61a 2.54a 2.46b
ORAC
(μmolTrolox g−1) 5 25 20 14 55 41 39 230 192
(μmolTrolox g1 DM) 44a 43a 35b 85a 82a 60b 424a 417a 346b

nd : not determined.

corroborate previous findings by Cisse, Vaillant, et al. (2009) on the tion, of colored compounds that interfere by increasing absorbance
thermal degradation of anthocyanins from roselle extract (Cisse, during the spectrophotometric determination of phenolics, as thermal
Vaillant, et al., 2009). However, it must be taken into account that high treatments often strongly affect juice color (Ding et al., 2008;
temperature treatments tend to lead to significantly decreased Rattanathanalerk, Chiewchan, & Srichumpoung, 2005).
stability of anthocyanins during storage (Cisse, Vaillant, Kane, Ndiaye, This hypothesis tends to be confirmed by the antioxidant values
& Domier 2011). assessed in the ORAC assay. The VE concentrate presented lower
For total phenolic content, losses during thermal concentration ORAC values (P b 0.05) than either the initial juice or OE concentrate. A
were inconsistent. In roselle extract, polyphenols were even higher in real increase in phenolic compounds would also have increased the
the VE concentrate than in the initial roselle extract. This surprising antioxidant potential. The OE concentrates and initial products
result was probably due to the formation, during thermal concentra- showed no significant differences in their ORAC values. For initial

Table 3
Comparison of different aroma compounds between the initial roselle extract, the concentrate obtained by osmotic evaporation (OE) and concentrate under vacuum (VE). Rt:
retention time; and IKC, IKL: calculated and literature Kovats indices.

Linear retention Peak areai / Peak areais Losses (%)

Aroma compounds Rt (min) IKC IKL Initial OE VE OE VE

Aldehyde Hexanal 6.18 1076 1084 8.2 (0.24) 7.0 (0.49) 2.3 (0.05) 15 72
Heptanal 9.25 1179 1174 2.4 (0.10) 2.3 (0.08) 0.9 (0.05) 4 63
Octanal 13.04 1283 1280 4.0 (0.08) 4.0 (0.20) 2.4 (0.40) 1 40
Nonanal 17.16 1387 1385 13.0 (0.29) 12.4 (0.20) 6.3 (0.04) 4 52
Furfural 19.72 1451 1455 39.7 (1.49) 32.5 (0.64) 12.6 (0.38) 18 68
Decanal 21.35 1491 1484 4.4 (0.10) 4.3 (0.54) 2.7 (0.24) 2 39
Benzaldehyde 21.92 1505 1495 4.1 (0.15) 3.8 (0.18) 3.2 (0.25) 7 21
2,4 dimethyl benzaldehyde 32.56 1788 – 18.7 (0.72) 17.5 (0.79) 7.5 (0.93) 7 60
Alcohols Octanol 23.95 1557 1553 9.1 (0.20) 6.6 (0.31) 1.9 (0.16) 27 79
Nonanol 27.84 1659 – 2.0 (0.20) 1.5 (0.07) 0.5 (0.06) 26 76
Eugenol 44.27 – 6.8 (0.14) 6.4 (0.83) 1.6 (0.54) 6 77
Others 2 Pentyl Furan 10.89 1226 1240 2.5 (0.05) 2.0 (0.12) 1.4 (0.03) 19 45
Methyl 6 heptadienon 3.5 24.85 1580 – 3.2 (0.07) 2.9 (0.09) 0.9 (0.14) 7 69
Trans decene 2 al 26.11 1613 1611 4.7 (0.45) 4.3 (0.27) 2.9 (0.05) 9 39
Linalyl propionate 29.00 1691 1698 6.6 (0.23) 5.3 (0.53) 2.0 (0.21) 20 63
Aromatic distance 8.1 32.5
358 M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360

Table 4
Comparison of different aroma compounds between the initial grape juice, the concentrate obtained by osmotic evaporation (OE) and concentrate under vacuum (VE). Rt: retention
time; and IKC, IKL: calculated and literature Kovats indices.

Linear retention Peak areai / Peak areais Losses (%)

Aroma compounds Rt (min) IKC IKL Initial OE VE OE VE

Aldehyde Octanal 13.12 1286 1280 2.0 (0.23) 1.8 (0.31) 1.4 (0.14) 12 32
Nonanal 17.25 1390 1385 7.8 (1.55) 6.1 (0.35) 4.4 (0.34) 22 43
Furfural 19.82 1453 1455 5.3 (0.12) 4.9 (0.05) 3.2 (1.00) 6 39
2 Phenyl ethanal 26.59 1626 1625 5.1 (0.18) 4.3 (0.04) 2.6 (0.31) 16 48
2 4 dimethyl benzaldehyde 32.69 1792 – 19.5 (0.29) 14.4 (0.51) 1.3 (0.08) 26 93
Alcohols Hexanol 15.89 1356 1360 8.9 (0.24) 6.2 (0.22) 0 31 100
Ethanol 3.57 934 929 19.2 (1.71) 14.9 (0.11) 0 23 100
2 Phenyl ethanol 36.43 1901 – 3.2 (0.19) 2.9 (0.15) 0 8 100
3 Methyl 1 butanol 10.29 1211 1205 7.3 (0.43) 55 (0.16) 0.7 (0.12) 24 90
Others Linalool 23.60 1549 1537 12.8 (0.17) 10.7 (0.01) 1.2 (0.03) 17 90
Linalyl propionate 29.11 1693 1698 7.5 (0.21) 5.5 (0.23) 0.8 (0.20) 26 90
Aromatic distance 8.3 32.2

products, the ORAC values obtained for grape juice and roselle extract grape juice, and apple juice, but between 21 and 79%, 32 and 100%, and
agreed with those of previous studies (Juliani et al., 2009; Lam, 71 and 100%, respectively, when vacuum evaporation was used.
Proctor, Howard, & Cho, 2005; Xu, Yuan, & Chang, 2007). Indeed, the losses reported for all aroma compounds were higher
Another difference between OE and VE concentrates refers to in the VE concentrates than in the OE concentrates, regardless of the
viscosity. For roselle extract, the viscosity of its VE concentrate was juice or extract tested. Because of the significant losses of aroma
two times higher than that of the OE concentrate. This could be compounds, thermal concentration by vacuum evaporation must be
explained by the gelatinization of pectin in a more acidic medium (pH coupled with an aroma recovery system, even though this system is
2.3). It also shows that polymerization reactions may have been often expensive and only partly efficient.
enhanced by a higher temperature (80 °C). Although the losses reported for osmotic evaporation were
Consequently, the physico-chemical properties of apple juice, grape relatively low, they were still significant. Other authors (Shaw et al.,
juice, and roselle extract were not significantly modified by osmotic 2001) found a global loss of volatile components of about 32% and
evaporation, whereas this was not the case for vacuum evaporation. 39%, respectively, in orange and passion fruit juices during osmotic
evaporation, using the same membrane type but in a system where
brine was continuously reconstituted by thermal evaporation.
3.3. Aromatic quality of concentrates In our case, however, losses seemed less important for each
product, as not all volatiles were equally affected because the
Tables 3, 4, and 5 report the relative losses of main aroma transmembrane transfer fluxes were different. In fact, during osmotic
compounds for roselle extract, grape juice, and apple juice, respectively, evaporation, mass transfer depended on (a) the initial concentration
during osmotic evaporation and thermal concentration under vacuum. of the compound in the juice, (b) its relative volatility, (c) its
For both processes, losses of volatile aroma compounds are observed. diffusivity in liquid phases and, finally (d), its diffusivity in air
Even so, losses are much lower for osmotic evaporation, being between entrapped within the membrane pores (Ali, 2004; Ali et al., 2003). The
1 and 27%, 6 and 31%, and 12 and 46%, respectively, for roselle extract, transmembrane transfers of aroma compounds also depended

Table 5
Comparison of different aroma compounds between the initial apple juice, the concentrate obtained by osmotic evaporation (OE) and concentrate under vacuum (VE). Rt: retention
time; and IKC, IKL: calculated and literature Kovats indices.

Linear retention Peak areai / Peak areais Losses (%)

Aroma compounds Rt (min) IKC IKL Initial COE CVE OE VE

Aldehydes Hexanal 6.18 1076 1084 17.0 (0.25) 12.8 (0.13) 1.2 (0.11) 25 93
1 trans 2 hexenal 10.34 1212 1220 9.6 (0.43) 7.5 (0.07) 1.3 (0.16) 21 87
2,4 Dimethyl benzaldehyde 32.57 1788 – 20.6 (0.97) 18.1 (0.09) 5.9 (0.29) 12 71
Alcohols Ethanol 3.53 931 929 9.1 (0.41) 5.5 (0.10) 0 40 100
Hexanol 15.82 1354 1360 183 (6.41) 112 (0.63) 0.3 (0.05) 39 100
Hex-2-enol 17.87 1405 1407 16.7 (0.48) 10.8 (0.07) 0 36 100
1 Octanol 23.95 1555 1553 5.1 (0.27) 3.1 (0.06) 0 39 100
1,2 methyl butanol 10.18 1208 1208 13.7 (0.48) 9.8 (0.15) 0 28 100
Esters Isoamyl acetate 7.29 1117 1117 19.8 (0.34) 10.6 (0.11) 0 46 100
Butyl formiate 8.12 1144 – 10.8 (0.36) 8.3 (0.16) 0 24 100
Butyl acetate 5.98 1067 1075 34.3 (1.08) 23.3 (0.45) 1.1 (0.17) 32 97
Hexyl acetate 12.50 1269 1270 19.2 (0.04) 12.8 (0.21) 0 33 100
2 hexenyl acetate 14.87 1330 1327 7.4 (0.13) 5.8 (0.03) 0 22 100
Ethyl butyrate 5.19 1032 1028 16.8 (0.05) 11.6 (0.11) 0 31 100
Methyl 1 ethyl 2 methyl butyrate 5.55 1048 1050 6.6 (0.10) 4.5 (0.07) 0 32 100
Pentyl methylbutyrate 20.92 1481 1488 5.2 (0.25) 4.1 (0.04) 0 21 100
Ethyl acetate 2.98 – 13.8 (0.27) 10.7 (0.03) 0 22 100
Others β-Damascenone 33.19 1807 1813 6.5 (0.60) 4.6 (0.09) 0.3 (0.03) 30 95
Aromatic distance 83 193
M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360 359

significantly on the characteristics of aromatic compounds and Cisse, M., Vaillant, F., Acosta, O., Dhuique-Mayer, C., & Dornier, M. (2009). Thermal
degradation kinetics of anthocyanins from blood orange, blackberry, and roselle
membrane materials, and on operating conditions. For instance, the using the arrhenius, eyring, and ball models. Journal of Agricultural and Food
aromatic “alcohol” class was globally more affected than other aroma Chemistry, 57(14), 6285−6291.
compound classes. Also, apple-juice aroma compounds generally Cisse, M., Vaillant, F., Kane, A., Ndiaye, O., & Domier, M. (2011). Anthocyanins and color
degradation kinetics of roselle extracts during storage. Journal of the Science of Food
seemed more sensitive to transmembrane transfers than the aroma and Agriculture submitted.
compounds of either grape juice or roselle extract. Cisse, M., Vaillant, F., Perez, A., Dornier, M., & Reynes, M. (2005). The quality of orange
Also, some of the losses reported during osmotic evaporation could juice processed by coupling crossflow microfiltration and osmotic evaporation.
International Journal of Food Science & Technology, 40(1), 105−116.
be explained by absorption of aroma compounds into the membrane's Courel, M., Dornier, M., Herry, J. -M., Rios, G. M., & Reynes, M. (2000). Effect of operating
surface, as reported in previous studies (Ali et al., 2003; Cisse et al., conditions on water transport during the concentration of sucrose solutions by
2005). Indeed, by implementing preconditioning and saturating osmotic distillation. Journal of Membrane Science, 170(2), 281−289.
Courel, M., Dornier, M., Rios, G. M., & Reynes, M. (2000). Modelling of water transport in
membrane adsorption sites, aroma losses could be drastically reduced
osmotic distillation using asymmetric membrane. Journal of Membrane Science, 173
(Ali et al., 2003; Cisse et al., 2005). (1), 107−122.
However, when aromatic distances (AD) are compared, OE Dhuique-Mayer, C., Tbatou, M., Carail, M., Caris-Veyrat, C., Dornier, M., & Amiot, M. J.
concentrates are clearly close to the initial products, whereas the (2007). Thermal degradation of antioxidant micronutrients in citrus juice: Kinetics
and newly formed compounds. Journal of Agricultural and Food Chemistry, 55(10),
aromatic distances of VE concentrates from the initial products are 2 4209−4216.
to 4 times greater. Thermally formed compounds like Furans and Ding, Z., Liu, L., Yu, J., Ma, R., & Yang, Z. (2008). Concentrating the extract of traditional
hydroxymethylfurfural were also observed in VE concentrates. Chinese medicine by direct contact membrane distillation. Journal of Membrane
Science, 310(1–2), 539−549.
Drosos, J. C., Viola-Rhenals, M., & Vivas-Reyes, R. (2010). Quantitative structure-
retention relationships of polycyclic aromatic hydrocarbons gas-chromatographic
4. Conclusions
retention indices. Journal of Chromatography A, 1217(26), 4411−4421.
Gancel, A. -L., Alter, P., Dhuique-Mayer, C., Ruales, J., & Vaillant, F. (2008). Identifying
The new osmotic pilot plant EURODIA/CIRAD, featuring a poly- carotenoids and phenolic compounds in Naranjilla (Solanum quitoense Lam. Var.
Puyo hybrid), an andean fruit. Journal of Agricultural and Food Chemistry, 56(24),
propylene, hollow-fiber membrane, presented good potential for
11890−11899.
concentrating fruit juices and plant extracts. Roselle extract could be George, S., Brat, P., Alter, P., & Amiot, M. J. (2005). Rapid determination of polyphenols
concentrated, at temperatures between 35 and 45 °C, by osmotic and vitamin c in plant-derived products. Journal of Agricultural and Food Chemistry,
evaporation to as much as 535 to 615 g TSS kg−1, with an average 53(5), 1370−1373.
Girard, B., & Fukumoto, L. R. (2000). Membrane processing of fruit juices and beverages:
evaporation flux of 1.1 to 1.5 kg h−1 m−2. The quality of the A review. Critical Reviews in Biotechnology, 20(2), 109−175.
concentrate obtained is similar to the initial roselle extract. Osmotic Hirunpanich, V., Utaipat, A., Morales, N. P., Bunyapraphatsara, N., Sato, H., Herunsale, A., &
evaporation did not modify contents of sugars, acids, polyphenols, or Suthisisang, C. (2006). Hypocholesterolemic and antioxidant effects of aqueous
extracts from the dried calyx of Hibiscus sabdariffa L. in hypercholesterolemic rats.
anthocyanins. Despite relative losses of aroma compounds, aromatic Journal of Ethnopharmacology, 103(2), 252−260.
quality was also preserved. Hogan, P. A., Canning, R. P., Peterson, P. A., Johnson, R. A., & Michaels, A. S. (1998). A new
Osmotic evaporation is a technology with an interesting potential option: Osmotic distillation. Chemical Engineering Progress, 94(7), 49−61.
Hongvaleerat, C., Cabral, L. M. C., Dornier, M., Reynes, M., & Ningsanond, S. (2008).
for concentrating, at low temperatures, tropical fruit juices and plant Concentration of pineapple juice by osmotic evaporation. Journal of Food
extracts, while maintaining their physico-chemical, biochemical, Engineering, 88(4), 548−552.
nutritional, and sensorial characteristics. With the industrial pilot Jiao, B., Cassano, A., & Drioli, E. (2004). Recent advances on membrane processes for
the concentration of fruit juices: A review. Journal of Food Engineering, 63(3),
plant used and under the operating conditions tested, the evaporation
303−324.
flow could be predicted, using the Reynolds number. Nevertheless, Juliani, H. R., Welch, C. R., Wu, Q., Diouf, B., Malainy, D., & Simon, J. E. (2009). Chemistry
before attempting industrial application, an economic study should be and quality of Hibiscus (Hibiscus sabdariffa) for developing the natural-product
industry in Senegal. Journal of Food Science, 74(2), S113−S121.
conducted to evaluate the impact of this technology on the cost of the
Lam, H. S., Proctor, A., Howard, L., & Cho, M. J. (2005). Rapid fruit extracts antioxidant
final product. capacity determination by Fourier transform infrared spectroscopy. Journal of Food
Science, 70(9), C545−C549.
Lee, J., Durst, R. W., & Wrolstad, R. E. (2005). Determination of total monomeric
Acknowledgment anthocyanin pigment content of fruit juices, beverages, natural colorants, and
wines by the pH differential method: Collaborative study. Journal of AOAC
International, 88, 1269−1278.
Authors wish to thank the French partner OSEO for its financial Lee, W. -C., Wang, C. -J., Chen, Y. -H., Hsu, J. -D., Cheng, S. -Y., Chen, H. -C., & Lee, H. -J.
support (oseo.fr). (2009). Polyphenol extracts from Hibiscus sabdariffa Linnaeus attenuate nephrop-
athy in experimental Type 1 diabetes. Journal of Agricultural and Food Chemistry, 57
(6), 2206−2210.
References Lin, T. -L., Lin, H. -H., Chen, C. -C., Lin, M. -C., Chou, M. -C., & Wang, C. -J. (2007). Hibiscus
sabdariffa extract reduces serum cholesterol in men and women. Nutrition Research,
Ali, F. 2004. Etude des transferts de composés d'arôme au cours de la concentration par 27(3), 140−145.
évaporation osmotique. Thèse de Doctorat en Génie des procédés, ENSIA. France Liu, J. -Y., Chen, C. -C., Wang, W. -H., Hsu, J. -D., Yang, M. -Y., & Wang, C. -J. (2006). The
227p. protective effects of Hibiscus sabdariffa extract on CCl4-induced liver fibrosis in rats.
Ali, F., Dornier, M., Duquenoy, A., & Reynes, M. (2003). Evaluating transfers of aroma Food and Chemical Toxicology, 44(3), 336−343.
compounds during the concentration of sucrose solutions by osmotic distillation in Odigie, I. P., Ettarh, R. R., & Adigun, S. A. (2003). Chronic administration of aqueous
a batch-type pilot plant. Journal of Food Engineering, 60(1), 1−8. extract of Hibiscus sabdariffa attenuates hypertension and reverses cardiac
AOAC (1990). Fruits and fruits products. In K. Helrich (Ed.), Official methods of analysis of hypertrophy in 2K-1C hypertensive rats. Journal of Ethnopharmacology, 86(2–3),
the association of official analytical chemists international (pp. 910−928). Arlington, 181−185.
VA. Ou, B., Hampsch-Woodill, M., & Prior, R. L. (2001). Development and validation of
Banvolgyi, S., Horvath, S., Stefanovits-Banyai, E., Bekassy-Molnar, E., & Vatai, G. (2009). an improved oxygen radical absorbance capacity assay using fluorescein as
Integrated membrane process for blackcurrant (Ribes nigrum L.) juice concentration. the fluorescent probe. Journal of Agricultural and Food Chemistry, 49(10),
Third Membrane Science and Technology Conference of Visegrad Countries (PERMEA). 4619−4626.
Cassano, A., Drioli, E., Galaverna, G., Marchelli, R., Di Silvestro, G., & Cagnasso, P. (2003). Rattanathanalerk, M., Chiewchan, N., & Srichumpoung, W. (2005). Effect of thermal
Clarification and concentration of citrus and carrot juices by integrated membrane processing on the quality loss of pineapple juice. Journal of Food Engineering, 66(2),
processes. Journal of Food Engineering, 57(2), 153−163. 259−265.
Cassano, A., Figoli, A., Tagarelli, A., Sindona, G., & Drioli, E. (2006). Integrated membrane Rodrigues, R. B., Menezes, H. C., Cabral, L. M. C., Dornier, M., Rios, G. M., & Reynes, M.
process for the production of highly nutritional kiwifruit juice. Desalination, 189(1–3), (2004). Evaluation of reverse osmosis and osmotic evaporation to concen-
21−30. trate camu-camu juice (Myrciaria dubia). Journal of Food Engineering, 63(1),
Cassano, A., Jiao, B., & Drioli, E. (2004). Production of concentrated kiwifruit juice by 97−102.
integrated membrane process. Food Research International, 37(2), 139−148. Shaw, P. E., Lebrun, M., Dornier, M., Ducamp, M. N., Courel, M., & Reynes, M. (2001).
Cisse, M., Dornier, M., Sakho, M., Ndiaye, A., Reynes, M., & Sock, O. (2009). Le bissap Evaluation of concentrated orange and passionfruit juices prepared by osmotic
(Hibiscus sabdariffa L.): Composition et principales utilisations. Fruits, 64(3), evaporation. Lebensmittel-Wissenschaft Und-Technologie-Food Science and Technol-
179−193. ogy, 34(2), 60−65.
360 M. Cissé et al. / Innovative Food Science and Emerging Technologies 12 (2011) 352–360

Vaillant, F., Cisse, M., Chaverri, M., Perez, A., Dornier, M., Viquez, F., & Dhuique-Mayer, C. Vaillant, F., Millan, P., O'Brien, G., Dornier, M., Decloux, M., & Reynes, M. (1999).
(2005). Clarification and concentration of melon juice using membrane processes. Crossflow microfiltration of passion fruit juice after partial enzymatic liquefaction.
Innovative Food Science & Emerging Technologies, 6(2), 213−220. Journal of Food Engineering, 42(4), 215−224.
Vaillant, F., Jeanton, E., Dornier, M., O'Brien, G. M., Reynes, M., & Decloux, M. (2001). Xu, B., Yuan, S., & Chang, S. (2007). Comparative analyses of phenolic composition,
Concentration of passion fruit juice on an industrial pilot scale using osmotic antioxidant capacity, and color of cool season legumes and other selected food
evaporation. Journal of Food Engineering, 47(3), 195−202. legumes. Journal of Food Science, 72(2), S167−S177.

You might also like