Anda di halaman 1dari 34

This is an author generated postprint of the article: Chan, C.-H., Yusoff, R., & Ngoh, G.-C. (2014).

Modeling
and kinetics study of conventional and assisted batch solvent extraction. Chemical Engineering Research and
Design, 92(6), 1169-1186. doi: 10.1016/j.cherd.2013.10.001

Modeling and kinetics study of conventional and assisted batch


solvent extraction
Chung-Hung Chan a,*, Rozita Yusoff a, Gek-Cheng Ngoh a

University of Malaya, Department of Chemical Engineering, 50603 Kuala Lumpur, Malaysia.

ABSTRACT:
Batch solvent extraction techniques have been widely explored. On understanding its potential
significance, this article aims to review kinetics and modeling of various extraction techniques
which involve assisted means including microwave-assisted extraction (MAE), ultrasonicassisted extraction (UAE), pulse electric field (PEF) and high voltage electrical discharge
(HVED). This review includes a detailed discussion of the instrumental setup, extraction
mechanisms and their distinct advantages and disadvantages. Additionally, the impact of the
operating parameters on the extraction kinetics of the mentioned techniques are highlighted.
The review also covers the mathematical modeling based on Ficks law, chemical rate law and
empirical models. The established kinetic models of various extractions are also summarized
to facilitate better understanding.

Keywords: kinetic modeling, microwave-assisted extraction, ultrasonic-assisted extraction,


pulse electric field, high voltage electrical discharge

*Corresponding author. Tel: +6017 7680611; Fax: +603 79675319; Email address: ch_chan@um.edu.my
The published version is available on http://dx.doi.org/10.1016/j.cherd.2013.10.001

Contents
1. Introduction
2. Batch solvent extraction
3. Assisted solvent extraction techniques
3.1. Microwave-assisted extraction (MAE)
3.2. Ultrasonic-assisted extraction (UAE)
3.3. Electrically-assisted extraction (EAE)
4. Influences of operating parameters on extraction kinetics
4.1. General operating parameters of extraction
4.2. Specific parameters of assisted solvent extraction
5. General Mathematical modeling
5.1. Ficks law
5.1.1. Mass transfer in solid particles
5.1.2. Mass transfer in solid particles and the extraction solvent
5.1.3. Mass transfer with extraction temperature variation
5.1.4. Mass transfer with degradation of active compounds
5.1.5. Modeling with external mass transfer resistance
5.1.6. Modified Ficks law
5.2. Rate law
5.3. Empirical equations
6. Extraction models
7. Summary
Acknowledgement
References
Nomenclature

1. Introduction
Batch solvent extraction is commonly used to extract active compounds from plants. The
current food industry adopts this technique and focuses on extracting and recovering valuable
active compounds from different plants and waste residues such as grape pomace (Casazza et
al., 2010), orange peels (Inoue et al., 2010) and olive cake (Crcel et al., 2010). In conventional
solvent extraction, the solid sample is immersed in the solvent and the extract is collected after
the equilibrium extraction is reached. The efficiency of the solvent extraction can be enhanced
by employing microwaves (Amarni and Kadi, 2010; Gujar et al., 2010; Spigno and De Faveri,
2009; Xiao et al., 2012), ultrasounds (Crcel et al., 2010; Pan et al., 2012; Stanisavljevi et al.,
2007; Velickovic et al., 2008) and electrical fields and charges (Boussetta et al., 2011; Elbelghiti and Vorobiev, 2005; El Belghiti and Vorobiev, 2004; Moubarik et al., 2011) into the
extraction system. These assisted techniques offer unique advantages and features which are
suitable for specific extractions. These non-conventional extraction techniques are presumed
to replace conventional solvent techniques, which makes studying their kinetics mechanisms
and modeling essential. Such studies enable prediction of the extraction behavior which is
considered to be useful for scaling up of the process.

In general, mathematical modeling approach that applies to solvent extraction depends on the
operational mode. Many reviews on the kinetic modeling of continuous solvent extraction such
as supercritical fluid extraction have been reported (Diaz and Brignole, 2009; Oliveira et al.,
2011). Thus this has prompted the review on the kinetic modeling of batch solvent extraction
techniques including the assisted techniques. The fundamental approach to model the
extraction is through derivation of Ficks law (Chen and Chen, 2011; Cisset al., 2012; Franco
et al., 2007a; Franco et al., 2007b; Herode et al., 2003; Hojnik et al., 2008; RakotondramasyRabesiaka et al., 2010; Tsibranska et al., 2011; Wongkittipong et al., 2004; Xu et al., 2008).
Other mathematical approaches employed include rate law (Pan et al., 2012; Qu et al., 2010;
Rakotondramasy-Rabesiaka et al., 2007; Xiao et al., 2012), Pelegs empirical model (Boussetta
et al., 2011; Crcel et al., 2010) and other two-parametric empirical models. To have a better
grasp on the extraction models developed, the derivations of the models together with their
assumptions and applications are elucidated. The information and data presented in this article
are very useful for specific plant extraction processes.

2. Batch solvent extraction


The solvent extraction curve is typically comprised of a fast extraction step (washing stage)
and a slow extraction step (diffusion stage) as shown in Fig. 1 (Franco et al., 2007b; Perez et
al., 2011; So and Macdonald, 1986). The extraction mechanism starts when the solvent
molecules penetrate into the plant matrices, causing the cytoplasm layer to be exposed directly
to the solvent (Crossley and Aguilera, 2001).This facilitates the dissolution of the active
compounds into the solvent. In the beginning of the extraction process, the fast step
corresponds to a constant extraction rate (Rakotondramasy-Rabesiaka et al., 2009). At an
extremely fast rate, the period in this extraction step is difficult to determine (Franco et al.,
2007b). During the slow extraction step, active compounds diffuse from the interior of the plant
matrices and dissolve in the solvent. The extraction yield during this step is greatly dependent
on the cells that remain intact after the washing extraction step (Crossley and Aguilera, 2001).
In fact, the characteristics of washing and diffusion steps in the extraction can be determined
by the proportion of broken and intact cells after sample preparation, e.g. grinding, (So and

Extraction yield

Macdonald, 1986).

diffusion stage

Washing stage

time

Fig. 1: Typical extraction curve of batch solvent extraction of active compounds from plants

Sample grinding and soaking in solvent are commonly applied prior to extraction in order to
reduce the particle size of the sample for better diffusion mechanism (Tsibranska et al., 2011)
and to improve the penetration of the solvent into the plant structure (Gujar et al., 2010).
Improvement of the extraction kinetics can also be achieved using advanced pretreatment such

as steam explosion and instant controlled pressure drop (DIC). These pretreatment methods
fragment the sample forming microspores as it is decompressed through sudden release of high
steam pressure (Ben Amor and Allaf, 2009; Chen and Chen, 2011). This improves the washing
step of the extraction (Chen and Chen, 2011) and enhances the diffusion of the solute into the
solvent (Ben Amor and Allaf, 2009).

Various setups of conventional extraction systems are available. For instance, a basic setup
consists of a stirred vessel with a water bath for temperature control, as shown in Fig. 2. This
setup has been widely applied in the industry to provide convective bulk movement in the
solvent. This reduces the mass transfer barrier and enhances the extraction (Franco et al.,
2007a). In some applications, a condenser is attached to the top of the vessel to prevent
evaporation of solvent due to overheating during the extraction (Xu et al., 2008). The
drawbacks associated with the conventional extraction technique in terms of long extraction
time and high solvent consumption have triggered the development of new solvent extraction
techniques with assisted means to overcome these limitations.

Fig. 2: Schematic diagrams of conventional and assisted extraction systems

3. Assisted solvent extraction techniques


Recent development in solvent extraction techniques focuses on enhancing the conventional
techniques with the assistance of microwave heating, ultrasonic radiation, electrical fields and
charges. These processes can be incorporated into solvent extraction or as sample pretreatment
prior to extraction. Their mechanisms, advantages and drawbacks are discussed in the
following sections.

3.1. Microwave-assisted extraction (MAE)


One of the assisted means used for enhancing the conventional solvent extraction system is
MAE. In this system, the heating efficiency is improved by applying microwaves. The
microwave radiation penetrates into the targeted material and interacts with the polar molecules
through ionic conduction and dipole rotation (Sparr Eskilsson and Bjrklund, 2000) to generate
heat. The localized heating is based on the dielectric constant of the material (Mandal et al.,
2007). The effectiveness of MAE is attributed to its localized heating which increases the
internal pressure of the cells and consequently ruptures them (Zhou and Liu, 2006). The active
compounds then elute from the cells and get dissolved in the surrounding solvent. The
schematic diagram of MAE instrumental setup is illustrated in Fig. 2. Closed type, opened type
and other modified MAE setup can be found in the literature (Chan et al., 2011).

MAE has been proven to enhance the extraction yields and shorten the extraction time in many
extractions (Chen et al., 2007; Li et al., 2010; Yan et al., 2010). For instance, the concentration
of phenolic compounds extracted from black tea by MAE after 90 s was 43% higher than that
using traditional brewing after 210 s (Spigno and De Faveri, 2009). Besides, microwave
heating can significantly improve the washing step of extraction. As reported in the kinetic
study of MAE of oils from olive cake (Amarni and Kadi, 2010), the rate constant of the washing
step of MAE was 17 times greater than that of the conventional extraction. This is probably
due to the rupture of the plant structure by microwave heating that enhanced the penetration of
the solvent into the interior structure of the plant sample (Gujar et al., 2010). The downside of
MAE can be referred to in the work (Chan et al., 2011).

3.2. Ultrasonic-assisted extraction (UAE)


Ultrasonic means is another technique which is capable of enhancing its mass transfer
mechanism in the extraction process. UAE provides stirring and thermal effects for the
extraction solvent, as well as structural effects for the solid sample. Ultrasound generates the

growth of bubbles inside liquids causing the cavitation phenomenon to occur, where the
cavitation bubbles implode asymmetrically near the solid surface (Leighton, 1998). This
phenomenon generates microjets in the direction of the solid surface and creates micro stirring
effects at the interface. It also reduces the effect of the boundary layers of the extraction and
results in enhanced mass transfer (Floros and Liang, 1994; Leighton, 1998). In addition, the
movement and implosions of the bubbles repeatedly squeeze and release the sample in what is
known as the Sponge effect (Jurez et al., 1999), can create micro channels in the sample
which improves the penetration of the solvent and provides larger contact area for mass transfer
(Jurez et al., 1999; Muralidhara et al., 1985). Another effect of ultrasound is attributed to its
thermo-acoustic influence which also has impact on the mass transfer resistance due to heating
(Mason and Lorimer, 2002). The instrumental setup for UAE is shown in Fig. 2. The extraction
temperature of the system is normally controlled through a water bath whereas other more
advanced setups of UAE can be referred to in the literature (Shirsath et al., 2012).

In UAE, the washing stage is enhanced due to cell destruction, improved solvent penetration
and mass transfer intensification (Vinatoru et al., 1999) but the diffusion stage is unffected
(Mili, 2013). As reported in UAE of oil from tobacco seeds (Stanisavljevi et al., 2007),
ultrasound improved the washing of oil from native seeds and some degree of milling effects
was observed in the seeds. However, the effect on ground seeds was insignificant
(Stanisavljevi et al., 2007). The stirring effect of UAE in reducing external mass transfer
resistance for diffusion was better than that provided by conventional extraction using
mechanical agitation as the former gave better extraction yields (Crcel et al., 2010). This was
proven by the results obtained from UAE of phenolic compounds from pomegranate peel in
which significant advantages were observed over conventional extraction techniques in terms
of extraction yields and extraction time (Pan et al., 2012). Nevertheless, UAE was less efficient
than conventional techniques in certain cases where oxidation and degradation of active
compounds occurred under prolonged sonication (Stanisavljevic et al., 2008; Stanisavljevi et
al., 2007). The interaction between the highly reactive hydroxyl radicals (resulted from
sonication) and the active compounds is responsible for the degradation (Karabegovic et al.,
2011). The key feature related to UAE is short extraction time though at times it has no positive
effect on the extraction yield as compared to conventional techniques (Stanisavljevi et al.,
2007; Velickovic et al., 2008).

3.3. Electrically-assisted extraction (EAE)


The kinetics of mass transfer can also be improved by exerting electrical effect during
extraction. There are two types of EAE techniques; high voltage electrical discharge (HVED)
and pulsed electric field (PEF). These two techniques operate in different mechanisms which
require different types of electrodes and applied voltage. As shown in Fig. 2, HVED requires
a high voltage generator and a needle electrode for arc discharge while PEF requires two
parallel plate electrodes for the induction of electric field.

The effectiveness of HVED on the extraction depends on the electrical breakdown between the
electrodes when high voltage is applied (Klimkin, 1990). This introduces energy directly into
the solvent-solid mixture through a plasma channel, which it is formed by a high voltage
electrical discharge generated between two submerged electrodes (Bogomaz et al., 1991). The
electrical breakdown creates high pressure shock waves and bubble cavitations which can
damage the cell structure and results in particle fragmentation (Gros et al., 2003; Touya et al.,
2006). Consequently, a better penetration of solvent into the solid particle enhances the mass
transfer mechanism. The performance of HVED in the extraction of polyphenols from wine
by-products was reported (Boussetta et al., 2011; Boussetta et al., 2012; Liu et al., 2011), and
the advantages of which were credited to its short treatment time (few ms) and low energy
consumption (10-50 kJ/kg) (Gros et al., 2003). With regards to the operational aspect, the
electrode gap distance is very critical for discharge formation and the optimum distance is very
much dependent on the nature of the extraction. An optimum electrode gap distance reduces
the amount of energy needed for plasma channel formation (Lang et al., 1998) and strengthens
the electric field, thus optimizes the discharge intensity (Sun et al., 1998). In addition, the pH
of the extraction system also affects the performance of HVED. The sample was found to be
relatively stable in acidic solution as compared to alkaline solutions as the latter can degrade
polyphenols when treated by HVED (Boussetta et al., 2011). This might have been caused by
the hydroxyl radicals which could have damaged the extracted compounds through oxidative
chemical reactions (Bogomaz et al., 1991; Chen et al., 2004; Ershov and Morozov, 2008).

Another electrically assisted technique is the pulsed electric field (PEF) which is based on the
electroporation phenomenon where it changes the permeability of cell membrane through a
potential difference across the membrane (Morales-Cid et al., 2010; Zimmermann et al., 1974).
During electroporation, molecular orientation takes place where the polar molecules align
themselves with the electric field and migrate to the membrane induced by the electric field

(Morales-Cid et al., 2010). The electrocompression exerted on the membrane ruptures the
membrane and creates pores on it (Soliva-Fortuny et al., 2009; Zimmermann et al., 1974). This
can result in a temporary (reversible) or permanent (irreversible) loss of membrane
permeability without having thermal alteration of the membrane (Morales-Cid et al., 2010;
Zimmermann et al., 1974). The permeability loss and the pore formation depend on the
induction of critical electric field strength and cell sizes in a range of 1-2 kV/cm for plant cells
sizes of 40-200 m (Heinz et al., 2001). PEF can substantially enhance the mass transfer in
biological tissues of different food plants with very low heating as well as improving both the
extraction yields during the washing step and the diffusion step (El-belghiti and Vorobiev,
2005).

The following section illustrates the influences of the important operating parameters on the
extraction kinetics of the various assisted solvent extraction techniques discussed above.

4. Influences of operating parameters on extraction kinetics


The effect of the general operating parameters such as the extraction solvent, solvent to feed
ratio, sample particle size and temperature are considered crucial for both conventional and
non conventional techniques. These parameters as well as some specific parameters for certain
assisted solvent extraction techniques such as microwave power, intensity of ultrasonic,
electrical field intensity and number of charges are discussed hereinafter to provide a better
insight of the extraction kinetics.

4.1. General operating parameters of extraction


The extraction solvent and its concentration play an important role in the extraction of active
compounds from plants. Different extraction solvents have different abilities to overcome the
energy barrier, which is also known as the activation energy of extraction (RakotondramasyRabesiaka et al., 2007; Spigno and De Faveri, 2009). This energy is required for the solvent to
penetrate into the interior of the plant cells. Aqueous organic solvents are usually employed in
most of the extraction. The mixture of ethanol and water is frequently used as an extraction
solvent in the extraction of active compounds from Fumaria officinalis L. (RakotondramasyRabesiaka et al., 2010), chestnut tree wood (Gironi and Piemonte, 2011) and grape seeds
(Bucic-Kojic et al., 2007). A suitable extraction solvent can enhance the washing step and

shorten the extraction time (Rakotondramasy-Rabesiaka et al., 2007). It can also improve the
diffusivity of the solute in the solvent and subsequently maximize the equilibrium extraction
yield (Xu et al., 2008).

Solvent to feed ratio is another important parameter which when applied correctly can decrease
the mass transfer barrier during the diffusion of active compounds and subsequently enhance
the extraction yield (Franco et al., 2007a; Qu et al., 2010). However, if the ratio is beyond the
optimum, the excess solvent does not have a significant effect on the equilibrium extraction
yields resulting in solvent wastage. It is worthy to note that the initial extraction rate during
the washing period is not significantly affected by the solvent to feed ratio (Herode et al.,
2003). On the other hand, extraction which is carried out at low solvent to feed ratio tends to
reach equilibrium much faster than those carried out at high solvent to feed ratio due to lower
equilibrium yields (Stanisavljevi et al., 2007).

The extraction kinetics is also greatly affected by the particle size of the plant sample. Slight
changes in particle size can significantly affect the extraction result. The extraction yield during
the washing step can be improved by smaller particle size of the sample (Qu et al., 2010) but it
becomes unfavorable when a small particle size sample leads to high extraction yields of
undesirable compounds (Cisset al., 2012). In terms of extraction kinetics, smaller particle size
increases the diffusivity and enhances the mass transfer mechanism in the diffusion step. This
is due to larger contact surface area with the solvent and shorter average diffusion path of the
active compounds from the solid to the solvent (Ciss et al., 2012; Herode et al., 2003; Hojnik
et al., 2008). As a result, shorter extraction time is required. The classic example is the
extraction of antioxidants from pomegranate marc (Qu et al., 2010) whereby a remarkable
reduction of extraction time from 90 min to 2 min was achieved when the particle size was
reduced from 3.5 mm to 0.2 mm. One worth-noting point is that the particle size does not affect
the initial extraction rate in the washing step provided that the internal diffusion of the active
compounds is rate limiting (Herode et al., 2003). Even though the effect of particle size on
the extraction process is obvious, it depends on the geometry of the extraction sample. For a
sample with plate geometry such as leaves, the effect of the particle size is not significant since
the relevant dimension for the diffusion of active compounds is the thickness of the leaves
(Wongkittipong et al., 2004). The effect might turn significant only when the particle size of
the leaves are reduced below its thickness such as in powder form.

Conventionally, the extraction temperature is monitored as it affects the stability of the active
compounds and the extraction performance. For extracting thermally stable compounds, better
extraction yield at elevated temperatures was achieved in shorter extraction time (Cisset al.,
2012). This can be explained by the high temperature that increases the diffusivity of the solute
and decreases the energy barrier of the extraction (Ciss et al., 2012; RakotondramasyRabesiaka et al., 2007). The solvation power of the extraction solvent is also enhanced allowing
more active compounds to be dissolved in the solvent (Cisset al., 2012; RakotondramasyRabesiaka et al., 2007; Xu et al., 2008). Furthermore, the extraction rate at the washing step
was reported to increase with temperature which could probably be due to increasing solvation
power of the solvent (Rakotondramasy-Rabesiaka et al., 2007, 2010). On the contrary, for
thermal sensitive compounds, elevated extraction temperature exerts negative effect on
extraction and has to be avoided at all costs. The general trend exhibited by these compounds
is that the stability decreases with increasing extraction temperature (Cisset al., 2012). This
also signifies that the rate of degradation or decomposition of thermal sensitive compounds
depends on the extraction temperature (Xiao et al., 2012). Thus, the selection criteria of a
suitable extraction temperature should be based on both the extraction efficiency and thermal
stability of the active compounds.

4.2. Specific parameters of assisted solvent extraction


Having elucidated the pertinent operating parameters for solvent extraction, specific
parameters of certain assisted solvent extraction techniques will be discussed. These
parameters have equally important influence on the extraction efficiency.

For instance,

microwave power controls the rate of heating during MAE extraction. As reported, higher
microwave powers decrease the extraction time but increase the extraction yields (Chemat et
al., 2005; Mandal and Mandal, 2010; Xiao et al., 2008). In the stability study of flavonoids
during microwave radiation (Biesaga, 2011), it was found that increasing the microwave power
amplifies the degradation as higher microwave heating causes sudden rise in temperature.
Elevated temperatures may result in overheating and undesired solvent evaporation, leading to
poor yields especially for thermally sensitive extracts. Similarly, the selection of microwave
power and extraction time is crucial to prevent thermal degradation of heat sensitive active
compounds.

For UAE, the effect of the intensity of ultrasounds on the extraction yields is significant (Ma
et al., 2009). In general, increased intensity level enhances the extraction yields and speeds up
the extraction regardless of the mode of radiation (Pan et al., 2012). However, this phenomenon
is not applicable to all cases as ultrasonic extraction does not impose any significant
enhancement on the extraction in some applications (Mircea, 2001). The evidence can be
observed from the poor ultrasound effect on both the yield and the kinetics in the extraction of
oil from woad seeds (Isatis tinetoria), independent of sonication conditions (Romdhane and
Gourdon, 2002). The poor effects imposed might be due to the geometrical shape of the treated
sample. While ultrasound enhances the extraction of olive leaves as the laminar shaped leaves
probably had received the ultrasound homogeneously but this was not the case for the grape
stalks which are of irregular shape (Crcel et al., 2010).

The high voltage electrical discharge (HVED) in the context of electrical assisted extraction is
the key factor accountable for the extraction yields. The treatment time of HVED is based on
the number of discharges where the typical peak pulse voltage is at 40 kV (Boussetta et al.,
2011; Liu et al., 2011; Moubarik et al., 2011). Increasing the number of discharges accelerates
the EAE extraction and thus enhances the equilibrium yields. Both the extraction yields during
washing and diffusion stages is increased by increasing the number of discharges until it
reaches an optimum point (Boussetta et al., 2011; Moubarik et al., 2011). The washing and
diffusion coefficient are also improved by increasing the number of discharges.

Analogous to the previous observation on HVED, the equilibrium extraction yields of EAE at
the washing and diffusion stages, increases with increasing intensity of pulse electrical field
(PEF) regardless of the extraction rate that remain unchanged (El-belghiti and Vorobiev, 2005;
El Belghiti and Vorobiev, 2004; Moubarik et al., 2011). Similar effects were also reported
when the number of pulses were increased in the extraction of sugar from beets (El Belghiti
and Vorobiev, 2004). In this study, the yield of solute increased from 15 to 30% which
corresponds to an increase in the number of pulses from 100 to 250.

5. Mathematical modeling of batch solvent extraction


Many mathematical approaches can be used to model the extraction process. The modeling
equations are either theoretically derived or empirically formulated. The widely employed
Ficks law of diffusion, chemical kinetic equations and other two-parametric empirical
equations are applicable for most extraction curves. Therefore, this section concentrates on the
derivation and applications of the mathematical modeling of batch solvent extraction.
5.1. Ficks Law
The diffusion step in batch type extraction depends on two extraction mechanisms; internal
diffusion and external diffusion. The internal diffusion of active compounds, as explained in
Ficks law, is driven by the difference in concentration between the plant matrix and the bulk
solvent (Bird et al., 2006) as follows:
N D

dC
dx

(1)

where N is the mass flux of the solute, C is the concentration of the solute in the solid particle,
D is known as the diffusivity or diffusion coefficient for the solute in the solvent, and x is the
distance in the direction of the transfer. For external diffusion, the active compounds diffuse
from the external surface of the solid to the bulk liquid. The determination of the rate limiting
mechanism in the diffusion stage is important in kinetic modeling as it determines a suitable
mathematical approach to model the extraction. To ensure efficient extraction, the external
mass transfer resistance has to be minimized so that the rate of extraction is dependent only on
the internal diffusion of the active compounds. Diffusivity in Ficks law in Eq. (1) is an
important property that indicates the rate of mass transfer and it is useful for equipment design
(Perez et al., 2011). Most of the kinetic modeling in the literature investigates the diffusivity
or other mass transfer coefficients in solvent extraction.
Characterization of extraction can be performed via derivation of Ficks law with initial and
boundary conditions. The solution of the mass transfer problem can be obtained analytically or
numerically depending on the complexity of the equations involved. Some basic assumptions
(Crank, 1975) which can be used to simplify the mass transfer problem are as follows:
i.

Symmetrical and porous sample particles. The geometry of solid particles is assumed
to be spherical with radius of R or thin plate with half thickness of L.

ii.

The solid particle is assumed to be of a pseudo-homogeneous medium. The


concentration of the active compounds in the solid particle depends on time and radius,

r or thickness, x.
iii.

Uniform distribution of active compounds in the sample matrix.

iv.

Homogeneous mixing between solvent and plant sample particles. The concentration
of the solute in the solvent only depends on time.

v.

The mass transfer of active compounds from the solid is a diffusion phenomenon in
which the diffusion coefficient is independent of time.

vi.

Diffusion of the solute and other compounds are in parallel and no interaction between
them.

vii.

External mass transfer resistance is negligible. The concentration of the solute in the
solvent at the interior of the solid particle is equal to the concentration of the solute in
the bulk solvent.

5.1.1. Mass transfer in solid particles


One of the assumptions of the mass transfer is to treat the external mass transfer resistance as
negligible, which is crucial and it depends on the nature of the extraction. This assumption
simplifies most of the extraction problems. The extraction process model can thus be developed
by considering only the mass balance in a spherical solid particle as shown:
C
( DC )
t

(2)

Where, t is the extraction time. Considering solely the spherical geometry of particles with
radius r, the respective initial and boundary conditions can be written as follows:
t 0,

C C0

t 0,

C Ci 0

rR

(4)

t 0,

C
0
r

r0

(5)

(3)

where, C0 is the initial concentration of solute in the sample particle, Ci is the concentration of
solute at the interface of sample particle. With the assumption of negligible external mass
transfer resistance, the concentration at the particle interface will become zero as described in
Eq. (4). The ordinary differential equation (ODE) for both spherical and plate geometry of
sample can then be expressed as in Eq. (6) and Eq. (7) respectively (Crank, 1975):
Spherical:

2 R (1)n
Dn 2 2t
C C0
nr
1
sin
exp

Ci C0
R
R 2

r n 1 n

(6)

Plate:

4 (1)n
(2n 1) 2 2 Dt
C C0
(2n 1) x
1
cos
exp

Ci C0
2L
4 L2

n 0 2n 1

(7)

The mass of solute transferred from the sample particle at any time, M can be calculated by
integrating the concentration of solute over the radius or thickness of the particles in Eq. (6)
and Eq. (7) to obtain Eq. (8) and Eq. (9) respectively.
Spherical:
M
6
1 2
M

n
n1

Dn 2 2t
exp

R 2

(8)

Plate:
M
8
1 2
M

(2n 1) 2 2 Dt
exp

4 L2
n0 (2n 1)

(9)

where, M is the total amount of solute transferred after infinite time. After a certain time lapse
or usually after the washing stage, only the first term of the series remains significant (Spiro,
1988). Both Eq. (8) and Eq. (9) can then be reduced to the following form (Perez et al., 2011):

M (t )
1 Ae Bt
M

(10)

where A is the model constant, and B is the diffusion rate constant. Theoretically, B = 2D/r2
is for spherical particles and B = 2D/4L2 is for plate particles. The constant B might not
applicable under certain conditions when its value has to depend on the geometry of plant
samples. The expression in Eq. (10) can be further modified based on non-extracted fraction of
the solute in the sample particle, E. Rearrangement of which has the simplified version as
shown below (Chen and Chen, 2011):

E 1

M (t )
Ae Bt
M

ln E ln A Bt

(11)
(12)

Alternatively, the concentration of solute in the extraction solvent at any time, c can be
expressed by considering only the first term of Eq. (8) and (9) (Spiro, 1988) with the simplified
forms shown as follows:
Spherical:

9.87 Dt
ln 0.498
R2
c c

(13)

Plate:

c
9.87 Dt
ln 0.21
4 L2
c c

(14)

where c is the concentration of solute in extraction solvent after infinite time. By plotting Eq.
(13) or Eq. (14) using experimental extraction curve, two intersecting straight lines can be
drawn and the slope of the first line is steeper than the second. The intersection between the
lines is the transition point. This denotes the point where the extraction changes its phase from
the washing step to the diffusion step (Kandiah and Spiro, 1990; Spiro et al., 1989). To achieve
better modeling results, Osburn and Katz (1944) suggested that the modeling of the extraction
process should consider both the washing and diffusion steps in Eq. (13) or Eq. (14) to yield
the following equations:
Spherical:
2
2

c c
6
D1t

D2t

2 f1 exp

f
exp

2
2
2
c


R
R

(15)

Plate:
2
2

c c
8
D1t

D2t

2 f1 exp

f
exp

2
2
2
c


4L

4L

(16)

where f1 and f2 are fractions of the solute extracted from the washing and diffusion stages with
diffusion coefficient of D1 and D2, respectively. The parameters D2 and f2 can be determined
from the slope and the intersection points of Eq. (13) or Eq. (14) as the second exponential
term is significant for the second stage of the extraction. In the early stage of extraction, the
second exponential term is close to unity thus D1 and f1 can be determined.

5.1.2. Mass transfer in solid particles and the extraction solvent


The models developed in the previous section can be theoretically obtained by solving the mass
transfer in solid particles. To make the models more realistic, mass balance in the solvent
should be taken into account. From the mass balance in solid particles shown previously in Eq
(2), the mass balance in the solvent can be expressed as shown in Eq (17):

VL

dc
J (t )
dt

(17)

where, VL is the volume of solvent used in the extraction. The initial extraction conditions and
the boundary conditions are presented in Eq. (18 &19) and Eq. (20-22) respectively:

t 0,

C C0

t 0

c0

(18)
(19)

At center, r = 0

r 0,

C
0
r

(20)

At interface, r = R
Diffusive flux from the solid particle (Cisset al., 2012; Wongkittipong et al., 2004; Xu et al.,
2008):

J (t ) DAs

C
r

(21)

Incoming flux in solvent (Tsibranska et al., 2011; Wongkittipong et al., 2004):

J (t ) Vs

dC
dt

(22)

where As is the specific area of the solid particle and Vs is the volume of plant sample. There
are two different boundary conditions at the solid interface which need to be considered;
namely the diffusive flux for the binary mixture in Eq. (21) and the incoming flux of the solvent
in Eq. (22). The same assumption of negligible external mass transfer resistance applies for
both Eq. (21) and Eq. (22). These equations can be used interchangeably or together depending
on the number of independent variables that need to be solved (Wongkittipong et al., 2004).
For example, in the modeling of extraction process by Wongkittipong et al. (2004), Eq. (21)
and Eq. (22) were used to solve the mass transfer problem that involved solid particles with
both cylindrical and plate geometry. This modeling approach is also applicable for other
geometry when the geometry shape factor is known (Wongkittipong et al., 2004).

5.1.3. Mass transfer with extraction temperature variation


The modeling approach presented so far is confined to specific extraction temperature. To
investigate the influence of temperature on the extraction process, Arrhenius equation shown
in Eq. (23) can be used to describe the effect of temperature on the diffusivity of the system.
Ea
D A' exp

RT

(23)

where, A is the pre-exponential factor and Ea is the activation energy of the Arrhenius model.
By comparing the reference diffusivity value, Dref at the reference temperature, Tref, the

diffusivity D at the extraction temperature T can be expressed as follows:

E 1
D
A' exp Ea RT
1

exp a

Dref
A' exp Ea RTref
R Tref T

(24)

By substituting this temperature related expression into the modeling system previously
discussed, the influence of temperature on the extraction process can be determined (Xu et al.,
2008).

5.1.4. Mass transfer with degradation of active compounds


When the extraction involves thermally sensitive compounds such as vitamins, the extraction
profile is driven by two processes namely; the diffusion of active compounds from plant
samples and the thermal degradation of active compounds in the extraction solvent (Xiao et al.,
2012):
k

D
deg
plant
Solvent

Decomposit ion

To improve the accuracy of the modeling, degradation terms can be added as first order rate
equations in the mass balance expressions for solid particles and solvent as shown in Eq. (25)
and Eq. (26) respectively (Cisset al., 2012).
Mass balance in solid particle with degradation term:
C
( DC ) k degC
t

(25)

Mass balance in solvent with degradation term:

VL

dc
J (t ) k degcV L
dt

(26)

where, kdeg describes the degradation constant of the extraction which can be related to
Arrhenius equations as shown in Eq.(27):

Ea
kdeg k exp

RT

(27)

where, k is the degradation rate constant for the active compound. The parameters in Eq. (27),
i.e. k and Ea, and the initial and boundary conditions in Eq. (18-22) are required to solve the
thermal degradation associated with the mass transfer problem.

5.1.5. Mass transfer with external mass transfer resistance


When the external mass transfer resistance in the extraction system becomes significant, the
convective mass transfer coefficient should be considered. A dimensionless form of Eq. (2) is
required to model this particular condition (Franco et al., 2007b).
2Y
r

Y

t

(28)

where the dimensionless groups for radius, extraction time and the yield can be defined as in
Eq. (29 31):
Dimensionless radius

r
R

(29)

Dimensionless extraction time

D t

(30)

R2

Dimensionless yield

(C Ce )
(C0 Ce )

(31)

where Ce is defined as the concentration of solute which remains in the sample particle after
infinite extraction time. The initial and boundary conditions are similar to those in Eq. (3) and
Eq. (5), only the boundary conditions at the interface are different and defined as follows:

t 0,

Y
r

kc R
k R c c

Y c
D
D c

(32)

where, kc is the convective mass transfer coefficient. This mass transfer problem was reported
by Walas (1991) and the suggested analytical solution is shown below:
_ _

Y (r , t )

n 1

_
Bn
sin(n 2 t )
r

(33)

while Bn can be determined from Eq. (34):

4 sin(n ) n cos( n ) n
Bn
2 n (sin(n )) 2
n

(34)

where n is the eigenvalues of the function given in Eq. (35):

cot g ( )

kc R
1
D

(35)

The average non-extracted fraction of the solute in the sample particle can be obtained from
Eq. (33) as follows:
_

_ _

Y Y (r ,. t ) d r
0
_

Y (t )

(36)

(0.8415 n ) exp( 2n t )

n 1

(37)

These series can be truncated to the first five terms to estimate the non-extracted oil fraction
with minor error.
5.1.6. Modified Ficks law
The models presented earlier are derived fundamentally and tre suitable to be used in scaling
up study and equipment design as the parameters involved in the equations have real physical
meanings. More simplified models like the film theory (Stankovicet al., 1994; Veljkovicand
Milenovic, 2002) and unsteady state diffusion through plant material (Ponomaryov, 1976;
Velickovic et al., 2006) can be adopted to describe the washing step and the diffusion step in
the extraction process. These two-parametric equations are derived from Ficks law and are
expressed in Eq. (38) and (39) respectively:
Film theory:

c
1 (1 b)e kt
c

(38)

Unsteady state diffusion theory:

C
(1 b ')e k 't
C0

(39)

where, b and b denote the coefficients for extraction kinetics in the washing step while k and
k are the coefficients for the diffusion step. To express Eq. (39) on the basis of the amount of
solute extracted in the extraction solvent, the equation can be modified into Eq. (40):

C0 C
(1 b ')e k 't
C0

(40)

where (C0-C) denotes the amount of solute dissolved in the extraction solvent. The modeling
equations involving Ficks law are commonly used in the modeling of solvent extraction as
they represent the fundamental theory for mass transfer.

In the next section, adaptation of another theory in the modeling of solvent extraction will be
discussed.

5.2. Rate law


Besides the expressions derived from Ficks Law, rate law had also been adapted in the
modeling of solvent extraction of active compounds from various plants in addition to being
employed to investigate the degradation rate of active compounds in the solution as discussed
previously (Xiao et al., 2012). Extraction models based on a second-order rate law are normally
applied in conventional and non-conventional extractions (Pan et al., 2012; Qu et al., 2010;
Rakotondramasy-Rabesiaka et al., 2007; Rakotondramasy-Rabesiaka et al., 2009). The rate of
dissolution of active compounds of a plant into the extraction solvent is given as follows:

dc
k1 c c 2
dt

(41)

where k1 is the second order extraction rate constant. Taking the initial and boundary conditions
as t = 0 to t and c = 0 to c, the integrated rate law can be obtained:

c2 k1t
c
1 c k1t

(42)

By linear transformation of Eq. (42), the rate constant k1 can be determined by fitting Eq. (43)
with experimental data. Subsequently Eq. (44) can be obtained from Eq. (43).

t
1
t

2
c k1 c c

(43)

c
1

t 1 k1c2 t c

(44)

c/t in Eq. (44) indicates the initial extraction rate, which can also be denoted by h, and can be
defined by Eq. (45) when the extraction time t approaches zero.

h k1c2

(45)

The concentration of solute in the extraction solvent at any time can then be described as:

t
1 h t c

(46)

5.3. Empirical equations


In the modeling of solvent extraction, various empirical models have been employed, either
developed from the fundamental models or adapted models as discussed above. The empirical
models are more suitable for extraction processes involving assisted means such as microwave,
ultrasound and electrical as they cannot be adequately described theoretically. The most
commonly used empirical model was proposed by So and Mcdonald (1986) and Patricelli et
al. (1979) which has the form shown in Eq. (47).

c cw 1 exp( k wt ) cd 1 exp( k d t )

(47)

where, cw and cd are the amounts of solute extracted in the solvent during the washing step and
the diffusion step, respectively. The amount of solute extracted can be expressed per mass of
sample used, or expressed in fraction by comparing with the equilibrium yield. kw and kd
represents the coefficients of extraction kinetics during the washing step and the diffusion step,
respectively. The empirical Eq. (47) resembles the model proposed by Osburn and Katz (1944)
previously shown in Eq. (15) and Eq. (16).
Pelegs model (Peleg, 1988), which was used to describe the sorption curves, was adapted for
the modeling of solvent extraction process as shown in Eq. (48):

c c0

t
K1 K 2t

(48)

where K1 is the model rate constant, K2 is the model capacity constant and C0 is usually equal
to zero. Both the Pelegs model and the second-order integrated rate law in Eq. (42) are
hyperbolic equations. Other empirical models proposed for solvent extraction include
Ponomaryov equation (Ponomaryov, 1976; Velickovic et al., 2006) and other two-parametric
empirical models such as parabolic diffusion model, power law and etc (Kitanovic et al., 2008).

6. Extraction models in batch solvent extraction


The kinetic models for various batch extraction techniques are tabulated in Table 1. This table
summarizes the model parameters with their associated operating conditions for the extraction.
There are two types of model parameters in the extraction namely; the rate of extraction either
for the washing or the diffusion steps, and the extraction capacity or equilibrium extraction
yield. The model parameters presented in Table 1 can be varied according to the plant sample
used, as the content of the active compounds may differ due to geographical location, weather

variation and soil conditions of the plantation. As a result, the kinetics as well as the extraction
capacity of the techniques at their designated operating conditions can be affected. As shown
in Table 1, the yield of the total extractive substances, total phenolic compounds, oil content
or the individual active compounds can be modeled by the corresponding equations given either
in terms of the concentration in the extraction solvent or the amount of extractives per sample
used.

From Table 1, it can be observed that the conventional extraction techniques are usually
modeled using Ficks law derivatives models as the model parameters contain physical
meanings which can be used for further interpretations. For example; Biot number (Bi) that
expresses the relative significance of internal and external mass transfer resistance (Seikova et
al., 2004). The theoretical based kinetic models are also suitable for scaling up purposes. On
the other hand, in the modeling of assisted solvent extraction techniques related to microwave,
ultrasonic, electrical field and charges; simplified models from modified Ficks law, empirical
models and rate law are preferred as these techniques implicate much more complicated mass
transfer problems.
Considering the sensitivity of the models coefficients with the change in the extraction
conditions, the increase or the decrease of the models coefficients would indicate certain
kinetic behavior of the extraction and the trend is dependent on their mathematical equations.
Most of the models coefficients increase to indicate enhanced kinetic (faster rate). However,
some coefficients, e.g. K1 and K2 of Peleg model, decreases to indicate a similar behavior.
Based on the tabulated data listed in Table 1, the effects of conventional operating parameters
on the extraction kinetics are in descending order of significance such as particle size of sample,
solvent to feed ratio and extraction temperature. Decrease in the particle size of sample and
increase in the solvent to feed ratio will enhance the diffusivity of Ficks law derivatives models
(D), the washing coefficient (b) and the diffusion coefficient (k) of modified Ficks law
models, and also the initial extraction rate (h) and extraction rate constant (k1) of rate law
models, predominantly. Once the parameters reached their optimum value, further increase in
the coefficients will not be significant. On the other hand, increase in the extraction temperature
also enhances the diffusivity (D). However, it may have opposite effect beyond certain
optimum temperature. This could probably be due to thermal degradation of active compounds.

Furthermore, the effects of specific parameters of assisted extraction on the models


coefficients as presented in Table 1 indicates that as microwave power (MAE) increases, the
washing and diffusion stages in terms of extraction yields (cw and cd) and kinetics coefficients
(kw and kd) generally will be enhanced provided there is no thermal degradation of active
compounds during the extraction. Furthermore, increase in the intensity of ultrasounds (UAE)
enhances the washing stage by increasing the initial extraction rate (h), and also the extraction
rate constant (k1) in the rate law models. In EAE, increase in the number of discharge (HVED)
improves both the washing and diffusion stages in terms of extraction yields (cw and cd) and
kinetics coefficients (kw and kd). However, increase in the intensity of electric field (PEF), can
only enhance the extraction yields (cw and cd).

Table 1: Kinetic models of batch solvent extraction


Authors

Extraction

Techniques

Perez et al.
(2011)

oil from
confectionery,
oilseed
(Helianthus
annuus) and
wild
(Helianthus
petiolaris)
sunflower
seeds.

conventional
extraction

Ben Amor and


Allaf (2009)

anthocyanins
from Malaysian
Roselle
(Hibiscus
sabdariffa)

DIC
pretreatment
followed by
conventional
extraction

Hojnik et al.
(2008)

lutein from
Marigold flower
petals

conventional
extraction

Franco et al.
(2007a)

oil from Rosa


rubiginosa

conventional
extraction

Operating
conditions
n-hexane,10 ml/g, 5
g sample, particle
sizes of 667 m
(confectionery), 624
m (oilseed) and 586
m (wild), stirring,
40-60 oC.

Equations

Model parameters

10
(spherical)

Confectionery:
A = 0.1361-0.1516 (50 oC) a
B = (1.19-1.66) x10-4 s-1 (50 oC) a
D = (1.34-1.87) x10-12 m2s-1
Oilseed:
A = 0.1148-0.1446 (40 oC) a
B = (2.09-5.10) x10-4 s-1 (60 oC) a
D = (2.06-5.03) x10-12 m2s-1
Wild:
A = 0.1798-0.21 (40 oC) a
B = (1.04-1.35) x10-4 s-1 (60 oC) a
D = (0.96-1.18) x10-12 m2s-1

water,100 ml/g, 2 g
DIC treated and
untreated sample,
particle sizes of 135
m, stirring, 100 oC

11 (plate)

DIC treated sample:


D = (4.62-6.11) x10-11 m2s-1
Untreated sample:
D = 4.19 x10-11 m2s-1

hexane, 10 ml/g
(total 500 ml
solvent), particle
sizes < 315 m,
stirring, 20-60 oC,
simultaneous
hydrolysis by adding
10% (w/v) alkali
solution at 7.5 ml/g
of sample mass into
extraction system to
obtain free lutein
ethanol,15-50 ml/g,
particle sizes < 600
m, stirring, 50 oC

15

20 oC:
D1= 6.128 x10-12 m2s-1
D2 = 0.011 x10-12 m2s-1
40 oC:
D1 = 2.175 x10-12 m2s-1
D2 = 0.018 x10-12 m2s-1
60 oC:
D1 = 1.503 x10-12 m2s-1
D2 = 0.011 x10-12 m2s-1

13

15 ml/g:
B = 0.0004 s-1
D = 0.61 x10-11 m2s-1
25 ml/g:
B = 0.0021 s-1
D = 3.19 x10-11 m2s-1
50 ml/g
B = 0.0046 s-1
D = 6.99 x10-11 m2s-1

Herode et al.
(2003)

carnosic acid
(CA), ursolic
acid (UA) and
oleanolic acid
(OA) from
Balm (Melissa
officinalis L.)
leaves

conventional
extraction

ethanol, 4-10 ml/g


(total 500 ml
solvent), particle
sizes of 200-400 m,
0-80 oC

16

CA, 20 oC, 4-10 ml/g:


D1 = (0.42-3.07) x10-11 m2s-1 (10 ml/g) a
D2 = (0.039-0.061) x10-11 m2s-1 (10
ml/g) a
UA, 20 oC, 4-10 ml/g:
D1 = (0.48-4.29) x10-11 m2s-1 (10 ml/g) a
D2 = (0.03-0.106) x10-11 m2s-1 (10 ml/g)
a

OA, 20 oC, 4-10 ml/g:


D1 = (0.28-2.59) x10-11 m2s-1 (8 ml/g) a
D2 = (0.045-0.119) x10-11 m2s-1 (6 ml/g)
a

CA, 4 ml/g, 0-80 oC:


D1 = (0.29-0.52) x10-11 m2s-1 (40 oC) a
D2 = (0.013-0.039) x10-11 m2s-1(40oC) a
UA, 4 ml/g, 0-80 oC:
D1 = (0.45-0.63) x10-11 m2s-1 b
D2 = (0.027-0.041) x10-11 m2s-1 (0 oC) a
OA, 4 ml/g, 0-80 oC:
D1 = (0.40-1.72) x10-11 m2s-1 (0 oC) a
D2 = (0.044-0.084) x10-11 m2s-1 (20oC) a
Tsibranska et al.
(2011)

total phenolic
compounds
from Sideritis
ssp. L.

conventional
extraction

80% Ethanol, 15
ml/g, particle size of
40 m, stirring, room
temperature

2, 17-20,
22

D = 1.5 x1012 m2s-1

Gujar et al.
(2010)

thymol from
seeds of
Trachyspermum
ammi

MAE

methanol, 30 ml/g, 1
g sample, 35-45 oC
(regulated by
microwave power at
0-300 W)

2, 17-20,
22

35 oC:
D = 1.835 x10-13 m2s-1
40 oC:
D = 2.46 x10-13 m2s-1
45 oC:
D = 3.24 x10-13 m2s-1

Xu et al. (2008)

isoflavones
from stem of
Pueraria lobata
(Willd.)

conventional
extraction

2, 17-21

D = 1.7 x10-11 m2s-1

Wongkittipong
et al. (2004)

andrographolide
from
Andrographis
paniculata

conventional
extraction
(Soxhlet)

50% n-butanol, 50
ml/g, 4 g sample,
particle size of 400800 m, stirring, 25
oC
60% ethanol, 50
ml/g, particle size of
600-800 m, 22-60
oC

2, 17-22

D = (8.43-52.1) x10-14 m2s-1 (60 oC) a

Cisset al.
(2012)

anthocyanins
from Hibiscus
sabdariffa

conventional
extraction

water, 25 ml/g,
particle sizes of 150
m, stirring, 25-90
oC
Ea = 61 kJ mol-1 c
k = 44200 s-1 c

18-21, 2527

D = (3.9-13.5) x10-11 m2 s-1 (90 oC) a

Franco et al.
(2007b)

oil from Rosa


rubiginosa

conventional
extraction

92% ethanol, 50
ml/g, particle sizes of
250, 350 and 750
m, stirring, 50 oC

28-37

250 m:
D = 5.2 x10-11 m2s-1
kR/D = 0.153
350 m:
D = 7.4 x10-11 m2s-1
kR/D = 0.131
750 m:
D = 30.0 x10-11 m2s-1
kR/D = 0.127

Velickovic et
al. (2006)

extractive
substances from
Salvia
officinalis L.
(SO) and Salvia
glutinosa L.
(SG) sage

Stanisavljevi et
al. (2007)

oil from
tobacco
(Nicotiana
tabacum L.)
seeds

Karabegovic et
al. (2011)

Pan et al. (2012)

Qu et al. (2010)

UAE

petroleum ether, 70%


ethanol and water as
solvent, 10 ml/g, 10 g
sample, total nominal
power of 3x50 W (40
kHz), 40 oC

38, 39

SO, petroleum ether:


b = 0.143
k = 1.14 x10-3 min-1
b' = 0.358
k' = 4.50 x10-3 min-1
SO, 70% ethanol:
b = 0.299
k = 2.21 x10-3 min-1
b' = 0.325
k' = 2.54 x10-3 min-1
SO, Water:
b = 0.251
k =1.86 x10-3 min-1
b' = 0.454
k' = 5.47 x10-3 min-1
SG, petroleum ether:
b = 0.144
k = 0.69 x10-3 min-1
b' = 0.50
k' = 5.09 x10-3 min-1
SG, 70% ethanol:
b = 0.254
k = 2.67 x10-3 min-1
b' = 0.276
k' = 3.05 x10-3 min-1
SG, water:
b = 0.233
k =1.51 x10-3 min-1
b' = 0.422
k' = 4.06 x10-3 min-1

UAE

n-hexane and
petroleum ether as
solvent, 3-10 ml/g, 5
g sample, total
nominal power 3x 50
W (40 kHz), 40 C

39

n-hexane:
b' = 0.5-0.6 (10 ml/g) a
k' = 4-7.8 x10-3 min-1 (3 ml/g) a
petroleum ether:
b' = 0.4-0.7 (10 ml/g) a
k' = 4-14 x10-3 min-1 (3 ml/g) a

flavonoids from
Artemisia
vulgaris (AV)
and Artemisia
campestris
(AC)

UAE

methanol, 10 ml/g,
10 g sample, total
nominal power 3x 50
W (40 kHz), 25 C

40

AV:
b'= 0.45
k'= 3.0 x10-3 min-1
AC:
b' = 0.4
k' = 3.8 x10-3 min-1

total phenolic
compounds
from
pomegranate
peel

UAE

water, 50ml/g,
operational modes:
1. Continuous mode,
intensity of 2.4-59.2
W/cm2 (20 kHz)
2. Pulse mode,
intensity of 59.2
W/cm2 (20 kHz),
pulse duration (s)
/interval (s) of 2/2,
5/5 and 5,15

41-46

Continuous mode:
h = 0.031-1.398 gL-1min-1 (59.2 W/cm2)

total phenolic
compounds
from
pomegranate
peel

conventional
extraction

water, 10-50 ml/g, 2


g sample, particle
sizes of 0.2-3.5 mm,
stirring, 25-90 oC
.

41-46

k1 = 0.012-0.185 Lg-1min-1 (59.2


W/cm2) a
Pulse mode, 2/2
h =1.128 gL-1min-1
k1 = 0.158 Lg-1min-1
Pulse mode, 5/5
h =1.456 gL-1min-1
k1 = 0.199 Lg-1min-1
Pulse mode, 5/15
h =0.498 gL-1min-1
k1 = 0.065 Lg-1min-1
50 ml/g, 0.2-3.5 mm, 25 oC:
h = 0.375-44.229 gL-1min-1 (0.2 mm) a
k1 = 0.115-7.685 Lg-1min-1 (0.2 mm) a
10-50 ml/g, 0.2 mm, 25 oC:
h = 12.967-50.905 gL-1min-1 (10 ml/g) a
k1 = 0.603-2.575 Lg-1min-1 (50 ml/g) a
50 ml/g, 0.2 mm, 25-90 oC:
h = 10.512-100.721 gL-1min-1 (95 oC) a
k1 = 1.948-6.314 Lg-1min-1 (95 oC) a

Rakotondramas
y-Rabesiaka et
al. (2009)

protopine from
Fumaria
officinalis
L.

conventional
extraction

0-94% aqueous,
6.25-100 ml/g (total
500 g solvent),
particle size of 0.40.5 mm, pH of 5.8,
stirring, 30 oC

41-46

Moubarik et al.
(2011)

extractive from
Foeniculum
vulgare

Pretreatments
by PEF, ED
and UAE
followed by
conventional
extraction

pretreatment:
1. PEF, water, 2
ml/g, 6.5g sample
(moisture 88-92%),
field intensity 0-430
V/cm (1000 Hz),
1000 pulses, pulse
repetition time of
10ms
2. ED, water, 2ml/g,
400 g sample, 40 kV,
pulse repetition time
of 10ms, 0-90 pulses
3. UAE, water, 2
ml/g, 30 g sample,
400 W/cm2 (40 kHz),
0-100 min
extraction:
Water, 2 ml/g,
stirring, 20 oC

47

Amarni and
Kadi (2010)

oil from olive


cake

MAE

hexane, 3 ml/g, 50 g
sample, stirring,
microwave power of
180-720 W

47

Bucic-Kojic et
al. (2007)

total
polyphenols
from
grape seeds

conventional
extraction

50% ethanol, 40
ml/g, 0.5 g sample,
particle sizes
of >0.63 mm, 0.630.4 mm, 0.4-0.16
mm, extraction
temperature at 25
oC, 50 oC and 80 oC

48

0-94% ethanol, 25 ml/g:


h = 3.8-15.9 mgL-1 min-1 (44 % ethanol)
a

k1 = 2.3-6.2 Lg-1 min-1 (0% ethanol) a


water, 6.25-100 ml/g:
h = 5.69-20.4 mgL-1min-1 (6.25 ml/g) a
k1 = 1.3-36.4 Lg-1min-1 (100 ml/g) a
44% ethanol, 6.25-100 ml/g:
h =7.28-30.3 mgL-1min-1 (6.25 ml/g) a
k1 = 0.5-17.5 Lg-1min-1 (100 ml/g) a
PEF:
cw = 0.43-0.54 (430 V/cm) a, d
cd = 0.31-0.43 (430 V/cm) a, d
kw= 0.17-0.22 b
kd = 0.011-0.015 b
ED:
cw = 0.43-0.55 (90 pulses) a, d
cd = 0.31-0.43 (90 pulses) a, d
kw= 0.214-0.286 (90 pulses) a
kd = 0.0144-0.0187 (90 pulses) a
UAE:
cw = 0.47-0.58 (100 min) a, d
cd = 0.27-0.39 (100 min) a, d
kw= 0.17-0.20(100 min) a
kd = 0.010-0.013 (100 min) a

cw = 4.02-4.22 g oil/ 100 g sample (720


W) a
cd = 0.58-1.07 g oil/ 100 g sample (720
W) a
kw = 7.54-250.14 (720 W) a
kd = 0.54-0.90 (540 W) a
>0.63 mm:
K1= 0.71-1.154 min (mg GAE/gdb)-1 (80
oC) e
K2 = 0.0393-0.0651 (mg GAE/gdb)-1 (80
oC) e
0.63-0.4 mm:
K1= 0.3056-0.4256min (mg GAE/gdb)-1
(80 C) e
K2 = 0.0375-0.0448 (mg GAE/gdb)-1 (80
C) e
0.4-0.16 mm:
K1= 0.0645-0.0869min/mg (50 C) e
K2 = 0.0175-0.0193 mg-1(50 C) e

The bracketed operating condition refers to the upper limit of the model parameter. b The effect of operating
condition on the model parameter is not significant. c the parameters used for computational obtained from other
studies. d Coefficient based on c/c. e The bracketed operating condition refers to lower limit of the model
parameter.

7. Summary
Various solvent extractions inclusive of kinetic modeling details have been compiled and
summarized in Table 2. This review gives an in-depth discussion on the extraction kinetics of
conventional and assisted solvent extraction techniques together with their pertinent operating
parameters. Mathematical modeling equations and kinetic models of extraction are also
presented. This review provides useful information pertaining to mathematical modeling of

various batch solvent extractions. The model parameters presented can also be utilized for the
comparative and scaling up studies.

Table 2: Summary of kinetic modeling of batch solvent extraction


Batch solvent extraction
Microwaveassisted
microwaveassisted
extraction
(MAE)
microwave
heating

Ultrasonicassisted

Electrically-assisted

ultrasonicassisted
extraction (UAE)

high voltage electrical


discharges (HVED);
pulsed electric field (PEF)

ultrasonic
radiation

Washing and
diffusion of active
compounds

Localized
heating of
microwave
builds internal
pressure to
rupture plant
cells

Cavitation
phenomenon
which provides
stirring and
structure effects
on plant structure

electric field, electrical


discharges
Electrical breakdown
creates high pressure
shock waves to damage
the cell structure (HVED);
Electroporation
phenomenon creates pore
on the membrane of plants
cells (PEF)

Improvements
on extraction
kinetic

improve
extraction rate in
washing and
diffusion step
resulted in short
extraction time

improve rate in
diffusion step,
enhance
equilibrium
extraction yield

enhance extraction yields


during washing and
diffusion steps (PEF),
accelerate extraction and
improve yields (HVED)

Important
parameters

solvent nature,
particle size of
sample and its
structure, solvent
to feed ratio,
temperature

microwave
power

ultrasound
intensity

intensity of Electric field


(PEF), Number of pulses
(PEF), Numbers of
discharge (HVED)

Modeling
equations
employed

Fick's law
derivations,
chemical rate law
and Peleg's model

Fick's law
derivations,
Patricelli's model

modified Fick's
law, chemical rate
law,

Patricelli's model

Conventional
Extraction
Technique

Features

Extraction
mechanism

batch extraction
with stirring and
temperature
control
temperature
control, agitation

ACKNOWLEDGEMENTS
This work was carried out under the Centre for Separation Science and Technology (CSST),
University of Malaya and financially supported through UMRG (RP002A-13AET) and PPP
(PV062/2011B) grants.

REFERENCES
Amarni, F., Kadi, H., 2010. Kinetics study of microwave-assisted solvent extraction of oil from olive cake using
hexane: Comparison with the conventional extraction. Innov. Food Sci. Emerg. Technol. 11(2), 322-327.
Ben Amor, B., Allaf, K., 2009. Impact of texturing using instant pressure drop treatment prior to solvent extraction
of anthocyanins from Malaysian Roselle (Hibiscus sabdariffa). Food Chem. 115(3), 820-825.
Biesaga, M., 2011. Influence of extraction methods on stability of flavonoids. J. Chromatogr. A 1218(18), 25052512.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2006. Transport Phenomena. Wiley.
Bogomaz, A.A., Goryachev, V.L., Remennyi, A.S., Rutberg, F.G., 1991. The Effectiveness of a pulsed electrical
discharge in decontaminating water. Sov. Tech. Phys. Lett. 17(6), 448- 449.
Boussetta, N., Vorobiev, E., Deloison, V., Pochez, F., Falcimaigne-Cordin, A., Lanoisell, J.L., 2011. Valorisation
of grape pomace by the extraction of phenolic antioxidants: Application of high voltage electrical discharges.
Food Chem. 128(2), 364-370.
Boussetta, N., Vorobiev, E., Le, L.H., Cordin-Falcimaigne, A., Lanoisell, J.L., 2012. Application of electrical
treatments in alcoholic solvent for polyphenols extraction from grape seeds. LWT - Food Sci. Technol. 46(1),
127-134.
Bucic-Kojic, A., Planinic, M., Tomas, S., Bilic, M., Velic, D., 2007. Study of solid-liquid extraction kinetics of
total polyphenols from grape seeds. J. Food Eng. 81(1), 236-242.
Crcel, J.A., Garca-Prez, J.V., Mulet, A., Rodrguez, L., Riera, E., 2010. Ultrasonically assistedantioxidant
extraction from grape stalks and olive leaves. Phys. Procedia 3(1), 147-152.
Casazza, A.A., Aliakbarian, B., Mantegna, S., Cravotto, G., Perego, P., 2010. Extraction of phenolics from Vitis
vinifera wastes using non-conventional techniques. J. Food Eng. 100(1), 50-55.
Chan, C.-H., Yusoff, R., Ngoh, G.-C., Kung, F.W.-L., 2011. Microwave-assisted extractions of active ingredients
from plants. J. Chromatogr. A 1218(37), 6213-6225.
Chemat, S., Ait-Amar, H., Lagha, A., Esveld, D.C., 2005. Microwave-assisted extraction kineticsof
from caraway seeds. Chem. Eng. Process. 44(12), 1320-1326.

terpenes

Chen, G., Chen, H., 2011. Extraction and deglycosylation of flavonoids from sumac fruits using steam explosion.
Food Chem. 126(4), 1934-1938.
Chen, Y.-S., Zhang, X.-S., Dai, Y.-C., Yuan, W.-K., 2004. Pulsed high-voltage discharge plasma for degradation
of phenol in aqueous solution. Sep. Purif. Technol. 34(13), 5-12.
Chen, Y., Xie, M.-Y., Gong, X.-F., 2007. Microwave-assisted extraction used for the isolation of total triterpenoid
saponins from Ganoderma atrum. J. Food Eng. 81(1), 162-170.
Ciss, M., Bohuon, P., Sambe, F., Kane, C., Sakho, M., Dornier, M., 2012. Aqueous extraction of anthocyanins
from Hibiscus sabdariffa: Experimental kinetics and modeling. J. Food Eng. 109(1), 16-21.
Crank, J., 1975. The mathematics of diffusion. Oxford University Press., Oxford.
Crossley, J.I., Aguilera, J.M., 2001. Modeling the effect of microstructure on food extraction. J. Food Process.
Eng. 24, 161177.
Diaz, M.S., Brignole, E.A., 2009. Modeling and optimization of supercritical fluid processes. J. Supercrit. Fluids
47(3), 611-618.

El-belghiti, K., Vorobiev, E., 2005. Modelling of Solute Aqueous Extraction from Carrots subjected to a Pulsed
Electric Field Pre-treatment. Biosyst. Eng. 90(3), 289-294.
El Belghiti, K., Vorobiev, E., 2004. Mass Transfer of Sugar from Beets Enhanced by Pulsed Electric Field. Food
Bioprod. Process. 82(3), 226-230.
Ershov, B., Morozov, P., 2008. Decomposition of ozone in water at pH 48. Russ. J. Appl. Chem. 81(11), 18951898.
Floros J.D., Liang, H., 1994. Acoustically assisted diffusion through membranes and biomaterials. Food Technol.
Biotech. 79, 84.
Franco, D., Pinelo, M., Sineiro, J., Nez, M.J., 2007a. Processing of Rosa rubiginosa: Extraction of oil and
antioxidant substances. Bioresource Technol. 98(18), 3506-3512.
Franco, D., Sineiro, J., Pinelo, M., Nez, M.J., 2007b. Ethanolic extraction of Rosa rubiginosa soluble
substances: Oil solubility equilibria and kinetic studies. J. Food Eng.79(1), 150- 157.
Gironi, F., Piemonte, V., 2011. Temperature and solvent effects on polyphenol extraction process from chestnut
tree wood. Chem. Eng. Res. Des. 89(7), 857-862.
Gros, C., Lanoisell, J.L., Vorobiev, E., 2003. Towards an Alternative Extraction Process for Linseed Oil. Chem.
Eng. Res. Des. 81(9), 1059-1065.
Gujar, J.G., Wagh, S.J., Gaikar, V.G., 2010. Experimental and modeling studies on microwave-assisted extraction
of thymol from seeds of Trachyspermum ammi (TA). Sep. Purif. Technol. 70(3), 257-264.
Heinz, V., Alvarez, I., Angersbach, A., Knorr, D., 2001. Preservation of liquid foods by high intensity pulsed
electric fieldsbasic concepts for process design. Trend Food Sci. Tech. 12(34), 103-111.
Herode, .S., Hadolin, M., kerget, M., Knez, ., 2003. Solvent extraction study of antioxidants from Balm
(Melissa officinalis L.) leaves. Food Chem. 80(2), 275-282.
Hojnik, M., Skerget, M., Knez, Z., 2008. Extraction of lutein from Marigold flower petals - Experimental kinetics
and modelling. LWT - Food Sci. Technol. 41(10), 2008-2016.
Inoue, T., Tsubaki, S., Ogawa, K., Onishi, K., Azuma, J.-i., 2010. Isolation of hesperidin from peels of thinned
Citrus unshiu fruits by microwave-assisted extraction. Food Chem. 123(2), 542-547.
Jurez, J.A.G., Corral, G.R., Moraleda, J.C.G., Yang, T.S., 1999. new high intensity ultrasonic technology for
food dehydration. Dry. Technol. 17, 587-608.
Kandiah, M., Spiro, M., 1990. Extraction of ginger rhizome: kinetic studies with supercritical carbon dioxide. Int.
J. Food Sci. Technol. 25, 328-338.
Karabegovic, I., Nikolova, M., Velickovic, D., Stojicevic, S., Veljkovic, V., Lazic, M., 2011. Comparison of
Antioxidant and Antimicrobial Activities of Methanolic Extracts of the Artemisia sp. Recovered by Different
Extraction Techniques. Chin. J. Chem. Eng. 19(3), 504-511.
Kitanovic, S., Milenovic, D., Veljkovic, V.B., 2008. Empirical kinetic models for the resinoid extraction from
aerial parts of St. John's wort (Hypericum perforatum L.). Biochem.Eng. J. 41(1), 1-11.
Klimkin, V.F., 1990. Mechanisms of electric breakdown of water from pointed anode in the nanosecond range.
Sov. Tech. Phys. Lett. 16, 146.
Lang, P.S., Ching, W.K., Willberg, D.M., Hoffmann, M.R., 1998. Oxidative degradation of 2, 4, 6-trinitrotoluene
by ozone in an electrohydraulic discharge reactor. Environ. Sci. Technol. 32, 3142-3148.
Leighton, T.G., 1998. The principles of cavitation, in: Povey, M.J.W., Mason, T.J. (Eds.), Ultrasound in Food
Processing. Chapman & Hall, London, pp. 151-182.

Li, J., Zu, Y.-G., Fu, Y.-J., Yang, Y.-C., Li, S.-M., Li, Z.-N., Wink, M., 2010. Optimization of microwave-assisted
extraction of triterpene saponins from defatted residue of yellow horn (Xanthoceras sorbifolia Bunge.) kernel
and evaluation of its antioxidant activity. Innov. Food Sci. Emerg. Technol. 11, 637.
Liu, D., Vorobiev, E., Savoire, R., Lanoisell, J.-L., 2011. Intensification of polyphenols extraction from grape
seeds by high voltage electrical discharges and extract concentration by dead-end ultrafiltration. Sep. Purif.
Technol. 81(2), 134-140.
Ma, Y.-Q., Chen, J.-C., Liu, D.-H., Ye, X.-Q., 2009. Simultaneous extraction of phenolic compounds of citrus
peel extracts: Effect of ultrasound. Ultrason. Sonochem. 16(1), 57-62.
Mandal, V., Mandal, S.C., 2010. Design and performance evaluation of a microwave based low carbon yielding
extraction technique for naturally occurring bioactive triterpenoid: Oleanolic acid. Biochem. Eng. J. 50(1-2), 6370.
Mandal, V., Mohan, Y., Hemalatha, S., 2007. Microwave assisted extraction - An innovative and promising
extraction tool for medicinal plant research. Pharmacog. Rev. 1(1), 7-18.
Mason, T.J., Lorimer, J.P., 2002. Applied Sonochemistry, The uses of power ultrasound in chemistry and
processing. Wiley-VCH, Weinheim.
Mili, P. S., Rajkovi, K. M., Stamenkovi, O. S., Veljkovi, V. B., 2013. Kinetic modeling and optimization of
maceration and ultrasound-extraction of resinoid from the aerial parts of white ladys bedstraw (Galium mollugo
L.). Ultrason. Sonochem. 20(1), 525-534.
Mircea, V., 2001. An overview of the ultrasonically assisted extraction of bioactive principles from herbs.
Ultrason. Sonochem. 8(3), 303-313.
Morales-Cid, G., Crdenas, S., Simonet, B.M., Valcrcel, M., 2010. Sample treatments improved by electric
fields. TrAC - Trend Anal. Chem. 29(2), 158-165.
Moubarik, A., El-Belghiti, K., Vorobiev, E., 2011. Kinetic model of solute aqueous extraction from Fennel
(Foeniculum vulgare) treated by pulsed electric field, electrical discharges and ultrasonic irradiations. Food
Bioprod. Process. 89(4), 356-361.
Muralidhara, H.S., Ensminger, D., Putnam, A., Drying Technology, 1985. Acoustic Dewatering and Drying (Low
and High Frequency): State of the Art Review. Dry. Technol. 3.
Oliveira, E.L.G., Silvestre, A.J.D., Silva, C.M., 2011. Review of kinetic models for supercritical fluid extraction.
Chem. Eng. Res. Des. 89(7), 1104-1117.
Osburn, J.O., Katz, D.L., 1944. Structure as a variable in the application of diffusion theory to extraction. Trans.
Am. Inst. Chem Eng. 40, 511-531.
Pan, Z., Qu, W., Ma, H., Atungulu, G.G., McHugh, T.H., 2012. Continuous and pulsed ultrasound-assisted
extractions of antioxidants from pomegranate peel. Ultrason. Sonochem. 19(2), 365-372.
Patricelli, A., Assogna, A., Casalaina, A., Emmi, E., Sodini, G., 1979. Fattori che influenzano l'estrazione dei
lipidi da semi decorticati di girasole. Riv. Ital Sostanze Gr. 56, 151-154.
Peleg, M., 1988. An empirical model for the description of moisture sorption curves. J. Food Sci. 53, 1216-1219.
Perez, E.E., Carelli, A.A., Crapiste, G.H., 2011. Temperature-dependent diffusion coefficient of oil from different
sunflower seeds during extraction with hexane. J. Food Eng. 105(1), 180-185.
Ponomaryov, V.D., 1976. Medicinal Herbs Extraction. Medicina, Moscow, in Russian.
Qu, W., Pan, Z., Ma, H., 2010. Extraction modeling and activities of antioxidants from pomegranate marc. J. Food
Eng. 99(1), 16-23.

Rakotondramasy-Rabesiaka, L., Havet, J.-L., Porte, C., Fauduet, H., 2007. Solid-liquid extraction of protopine
from Fumaria officinalis L.--Analysis determination, kinetic reaction and model building. Sep. Purif. Technol.
54(2), 253-261.
Rakotondramasy-Rabesiaka, L., Havet, J.-L., Porte, C., Fauduet, H., 2010. Estimation of effective diffusion and
transfer rate during the protopine extraction process from Fumaria officinalis L. Sep. Purif. Technol. 76(2), 126131.
Rakotondramasy-Rabesiaka, L., Havet, J.L., Porte, C., Fauduet, H., 2009. Solid-liquid extraction of protopine
from Fumaria officinalis L.--Kinetic modelling of influential parameters. Ind. Crop. Prod. 29(2-3), 516-523.
Romdhane, M., Gourdon, C., 2002. Investigation in solidliquid extraction: influence of ultrasound. Chem. Eng.
J. 87(1), 11-19.
Seikova, I., Simeonov, E., Ivanova, E., 2004. Protein leaching from tomato seedexperimental kinetics and
prediction of effective diffusivity. J. Food Eng 61, 165-171.
Shirsath, S.R., Sonawane, S.H., Gogate, P.R., 2012. Intensification of extraction of natural products using
ultrasonic irradiationsA review of current status. Chem. Eng. Process. 53(0), 10-23.
So, G.C., Macdonald, D.G., 1986. Kinetics of oil extraction from Canola (Rapeseed). Can. J. Chem. Eng. 64, 8086.
Soliva-Fortuny, R., Balasa, A., Knorr, D., Mart
n-Belloso, O., 2009. Effects of pulsed electric fields on bioactive
compounds in foods: a review. Trend Food Sci. Technol. 20(1112), 544-556.
Sparr Eskilsson, C., Bjrklund, E., 2000. Analytical-scale microwave-assisted extraction. J.
902(1), 227-250.

Chromatogr.

Spigno, G., De Faveri, D.M., 2009. Microwave-assisted extraction of tea phenols: A phenomenological study. J.
Food Eng. 93(2), 210-217.
Spiro, M., 1988. The rate of caffeine infusion from Kenyan arabica coffee beans. 12th Colloque Scientifique
International du Cafe, 260-264.
Spiro, M., Kandiah, M., Price, W., 1989. Extraction of ginger rhizome: kinetic studies with dichloromethane,
ethanol, 2-propanol and an acetone water mixture. Int. J. Food Sci. Technol. 25, 157-167.
Stanisavljevic, I., Stojicevic, S., Velickovic, D., Lazic, M., Veljkovic, V.S.S.T., 2008. Screening the antioxidant
and antimicrobial properties of the extracts from plantain (Plantago major L.) leaves. Sep. Sci. Technol. 43(13),
3652-3662.
Stanisavljevi, I.T., Lazi, M.L., Veljkovi, V.B., 2007. Ultrasonic extraction of oil from tobacco (Nicotiana
tabacum L.) seeds. Ultrason. Sonochem. 14(5), 646-652.
Stankovic, M. Z., Cakic, M. D., Cvetkovic, D. M., & Veljkovic, V. B. 1994. Kinetics of extraction of resinoids
from overground parts of sweet clover (Melilotus officinalis L). J. Serb. Chem. Soc., 10(39), 735-741.
Sun, B., Sato, M., Harano, A., Clements, J.S., 1998. Non-Uniform pulse discharge induced radical production in
distilled water. J. Electrostat. 43, 115-126.
Touya, G., Reess, T., Pecastaing, L., Gibert, A., Domens, P., 2006. Development of subsonic electrical discharges
in water and measurements of the associated pressure waves. J. Phys. D: Appl. Phys. 39(24), 5236-5244.
Tsibranska, I., Tylkowski, B., Kochanov, R., Alipieva, K., 2011. Extraction of biologically active compounds
from Sideritis ssp. L. Food Bioprod. Process. 89(4), 273-280.
Velickovic, D.T., Milenovic, D.M., Ristic, M.S., Veljkovic, V.B., 2006. Kinetics of ultrasonic extraction of
extractive substances from garden (Salvia officinalis L.) and glutinous (Salvia glutinosa L.) sage. Ultrason.
Sonochem. 13(2), 150-156.

Velickovic, D.T., Milenovic, D.M., Ristic, M.S., Veljkovic, V.B., 2008. Ultrasonic extraction of waste solid
residues from the Salvia sp. essential oil hydrodistillation. Biochem. Eng. J. 42(1), 97-104.
VeljkovicV., MilenovicD., 2002. Extraction of resinoids from St. Johns worth (Hypericum perforatum L.) II.
Modelling of extraction kinetics. Chem. Ind. (Belgrade) 56, 60-67, in Serbian.
Vinatoru, M., Toma, M., Mason, T., 1999. Ultrasonically assisted extraction of bioactive principles from plants
and their constituents. J. Adv. Sonochem. 5, 209-249.
Walas, S.M., 1991. Modeling with differential equations in chemical engineering. Butterworth-Heinemann,
Stoneham, MA, USA.
Wongkittipong, R., Prat, L., Damronglerd, S., Gourdon, C., 2004. Solid-liquid extraction of andrographolide from
plants--experimental study, kinetic reaction and model. Sep. Purif. Technol. 40(2), 147-154.
Xiao, W., Han, L., Shi, B., 2008. Microwave-assisted extraction of flavonoids from Radix Astragali. Sep. Purif.
Technol. 62(3), 614-618.
Xiao, X., Song, W., Wang, J., Li, G., 2012. Microwave-assisted extraction performed in low temperature and in
vacuo for the extraction of labile compounds in food samples. Anal. Chim. Acta 712(0), 85-93.
Xu, H.-N., Huang, W.-N., He, C.-H., 2008. Modeling for extraction of isoflavones from stem of Pueraria lobata
(Willd.) Ohwi using n-butanol/water two-phase solvent system. Sep. Purif. Technol. 62(3), 590-595.
Yan, M.-M., Liu, W., Fu, Y.-J., Zu, Y.-G., Chen, C.-Y., Luo, M., 2010. Optimisation of the microwave-assisted
extraction process for four main astragalosides in Radix Astragali. Food Chem. 119(4), 1663-1670.
Zhou, H.-Y., Liu, C.-Z., 2006. Microwave-assisted extraction of solanesol from tobacco leaves. J. Chromatogr. A
1129(1), 135-139.
Zimmermann, U., Pilwat, G., Riemann, F., 1974. Dielectic breakdown in cell membranes. Biophysic. J. 14, 881889.

NOMENCLATURE
A

model constant (dimensionless)

pre-exponential factor (dimensionless)

As

specific area of the solid particle (m2)

diffusion rate constant (s-1)

coefficient of extraction kinetic in the washing step (dimensionless)

coefficient of extraction kinetic in the washing step (dimensionless)

concentration of solute in the solid particle (g m-3)

concentration of solute in the extraction solvent at any time (g m -3)

C0

initial concentration of the solute in the sample particle (g m -3)

concentration of solute in the extraction solvent after infinite time (g m -3)

cd

amount of solute extracted in the solvent during diffusion step (dimensionless)

Ce

concentration of solute remains in the sample particle after infinite time (g m -3)

Ci

concentration of solute at the interface of the sample particle (g m -3)

cw

amount of solute extracted in the solvent during the washing step (dimensionless)

diffusion coefficient (m2 s-1)

D1

diffusion coefficient during the washing stage (m2 s-1)

D2

diffusion coefficient during the diffusion stage (m2 s-1)

Dref

diffusion coefficient at the reference temperature (m2 s-1)

non extracted fraction of solute in the sample particle (dimensionless)

Ea

activation energy

f1

fraction of solute extracted from the washing stage (dimensionless)

f2

fraction of solute extracted from the diffusion stage (dimensionless)

initial extraction rate (g L-1 min-1)

coefficient of extraction kinetic in the diffusion step (min -1)

coefficient of extraction kinetic in the diffusion step (min -1)

k1

second order extraction rate constant (L g-1 min-1)

K1

model rate constant (min mg-1)

K2

model capacity constant (mg-1)

degradation rate constant (s-1)

kc

convective mass transfer coefficient (m s-1)

kd

coefficient of extraction kinetics during the diffusion step (min -1)

kdeg

degradation constant of the extraction (s-1)

kw

coefficient of extraction kinetics during the washing step (min -1)

half of the thickness of the solid particle (m)

mass of solute transferred from the sample particle at any time (g)

total amount of solute transferred after infinite time (g)

mass flux of the solute (g m-2 s-1)

radial distance in the diffusion direction (m)

radius of the solid particle (m)

dimensionless radius (dimensionless)

time (s)

temperature (oC)

dimensionless time (dimensionless)

Tref

reference temperature (oC)

VL

volume of the solvent used in the extraction (m3)

Vs

volume of the plant sample (m3)

distance in the diffusion direction (m)

dimensionless yield (dimensionless)

average non-extracted fraction of solute in the sample particle (dimensionless)

Anda mungkin juga menyukai