You are on page 1of 26

This article was downloaded by: [University of Windsor]

On: 17 June 2013, At: 02:26


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Food Reviews International


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/lfri20

Use of Power Ultrasound to Improve


Extraction and Modify Phase Transitions
in Food Processing
a b c
Jayani Chandrapala , Christine M. Oliver , Sandra Kentish &
a
Muthupandian Ashokkumar
a
School of Chemistry, University of Melbourne, Melbourne, Victoria,
Australia
b
CSIRO Division of Food and Nutritional Sciences, Melbourne,
Victoria, Australia
c
Department of Chemical and Biomolecular Engineering, University
of Melbourne, Melbourne, Victoria, Australia
Accepted author version posted online: 29 May 2012.

To cite this article: Jayani Chandrapala , Christine M. Oliver , Sandra Kentish & Muthupandian
Ashokkumar (2013): Use of Power Ultrasound to Improve Extraction and Modify Phase Transitions in
Food Processing, Food Reviews International, 29:1, 67-91

To link to this article: http://dx.doi.org/10.1080/87559129.2012.692140

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Food Reviews International, 29:67–91, 2013
Copyright © Taylor & Francis Group, LLC
ISSN: 8755-9129 print / 1525-6103 online
DOI: 10.1080/87559129.2012.692140

Use of Power Ultrasound to Improve Extraction


and Modify Phase Transitions in Food Processing

JAYANI CHANDRAPALA1, CHRISTINE M. OLIVER2,


SANDRA KENTISH3, AND
MUTHUPANDIAN ASHOKKUMAR1
1
School of Chemistry, University of Melbourne, Melbourne, Victoria, Australia
2
CSIRO Division of Food and Nutritional Sciences, Melbourne, Victoria,
Australia
Downloaded by [University of Windsor] at 02:26 17 June 2013

3
Department of Chemical and Biomolecular Engineering, University of
Melbourne, Melbourne, Victoria, Australia

In recent years, the physical and chemical effects of ultrasound in liquid and solid media
have been extensively used in food processing applications. Ultrasound in liquids gener-
ates a number of physical forces. Vibration, pressure, and physical agitation are forces
that can be generated in the absence of acoustic cavitation. In addition to these phys-
ical forces, acoustic cavitation generates microjets, shear forces, shockwaves, radical
formation, and acoustic streaming. At lower frequencies (20–100 kHz), the physical
effects dominate. At intermediate frequencies (200–500 kHz), chemical effects (forma-
tion of highly reactive radicals within the cavitation bubbles) are more dominant, as
the number of active bubbles generated is higher. At higher frequencies (>1 MHz),
cavitation and the associated chemical effects are less likely and acoustic streaming
effects are dominant. There are a number of food processing applications where these
physical and chemical forces of ultrasound have been found to be effective. The present
review summarizes selected areas of food applications such as extraction, crystalliza-
tion, thawing, drying, and freezing where ultrasound is found to be beneficial in terms of
increasing efficiency, reducing time, and increasing the yields. The reason for choosing
these applications is that such areas are not critically reviewed in the existing literature.

Keywords Crystallization, drying, extraction, freezing, ultrasound

Introduction
Ultrasound refers to sound waves beyond the audible frequency range (in general,
>20 kHz). When ultrasound passes through a liquid medium, the interaction between
the ultrasonic waves, liquid, and dissolved gas leads to a phenomenon known as acoustic
cavitation, which occurs under specific experimental conditions. Acoustic cavitation can
be generated in liquids in the frequency range of 20 kHz to >1 MHz, above critical
power levels. Several review articles and textbook chapters describe acoustic cavitation
in detail.(1–4) Application of an acoustic field forces a free bubble to oscillate about its
equilibrium radius. During the contraction phase, the concentration of gas inside the bub-
ble increases and gas diffuses out of the bubble. Similarly, during the expansion phase,

Address correspondence to Muthupandian Ashokkumar, School of Chemistry, University of


Melbourne, Melbourne VIC 3010, Australia. E-mail: masho@unimelb.edu.au

67
68 Chandrapala et al.

the concentration decreases, and gas diffuses into the bubble. Since the diffusion rate is
proportional to the area, more gas enters during expansion than leaves during contraction;
consequently, over a complete cycle, there will be a net increase in the amount of gas inside
the bubble, a phenomenon called rectified diffusion. Rectified diffusion and bubble coa-
lescence lead to the growth of the bubbles towards a resonance size range. When bubbles
reach the resonance size range, they grow to a maximum size within one acoustic cycle
and violently collapse, generating very high temperature conditions within the collapsing
bubbles. Experimentally determined temperatures are about 2000–10,000 K, whereas the-
oretical predictions estimate temperatures of up to 100,000 K.(2,3) Owing to the generation
of such high temperatures within the cavitation bubbles, they are also referred to as “hot
spots.” Highly reactive radicals are generated within the cavitation bubbles. If water is
the medium, primary radicals such as OH• and H• are generated. In the presence of other
volatile solutes, for example, alcohols, a variety of radicals are generated.(4–6) In addition to
generating chemical reactions, cavitation also generates violent physical forces that include
Downloaded by [University of Windsor] at 02:26 17 June 2013

microjets, shear forces, shock waves, and turbulence.(2,3)

Ultrasound-Assisted Extraction
Ultrasound-assisted extraction (UAE) is an effective way of extracting a number of ana-
lytes from food and food-related materials. The efficiency of UAE is mainly attributed to
acoustic cavitation. The very high effective temperatures (which increase solubility and
diffusivity) and pressures (which facilitate penetration and mass transfer) at the interface
between a liquid medium subjected to ultrasonication and a solid matrix, combined with the
oxidative radicals generated during sonolysis, bring about high extractive power. In many
analytical situations, UAE is a simpler and more efficient alternative to conventional extrac-
tion (e.g., steam distillation, Soxhlet, maceration), and occasionally, modern extraction
techniques (e.g., microwave, supercritical fluid). As reported by Toma et al.,(7) the micro-
jetting and microstreaming effects attributed to acoustic cavitation cause disintegration
of solid materials and disruption of cell walls. This phenomenon involves an increase in
the contact between the solvent and the cell contents, and enhances mass transfer of the
cell contents from the material to the solvent.(8–10) In addition, as the size of the bubbles
generated during acoustic cavitation are very small relative to the total liquid volume, the
heat produced on bubble collapse is rapidly dissipated with no appreciable change in the
environmental conditions. In the case of dried substrates, ultrasound may be used to facil-
itate swelling and hydration,(7) and increase porosity.(8) Solution temperatures may also
affect the ultrasonic extraction efficiency. There are two opposing effects of the solution
temperature. Ultrasound (US)-assisted extraction may increase the extraction yield at low
temperatures, since acoustic cavitation is more efficient. On the other hand, solubility of
the extracted material may increase with an increase in solution temperature. There is no
systematic study available on the effect of solution temperature.
The main reported benefits of UAE over other extraction techniques include an
enhancement in extraction yield and rate, providing high throughput and reduced extrac-
tion time.(11) For example, Gómez-González et al.(12) reported that a solvent-assisted UAE
(power [P] = 450 W, frequency [f] = 20 kHz, 60% amplitude) of sugars from olive fruits,
stems, and leaves enabled complete isolation of sugars within 10 min, and provided an effi-
ciency similar to that achieved by a conventional maceration procedure in 24 h. As another
example, Mason and Zhao(13) enhanced the extraction of tea solids using ultrasound (f =
20 kHz, solid:liquid ratio of 1:30). The majority of the material was extracted by aque-
ous sonication of dried tea leaves within the first 10 min of ultrasound and increased the
Use of Ultrasound in Food Processing 69

yield up to 20% at 60 ◦ C, compared with water alone, approaching the efficiency of ther-
mal extraction at 100 ◦ C. In a subsequent study, independent researchers showed that UAE
(P = 250 W, f = 40 kHz, 60 ◦ C, 40 min) enhances both the chemical and sensory quality of
aqueous tea beverages.(14) Temperature is important when using ultrasound in tea extrac-
tion, since cavitation provides the driving force for enhanced extraction and indeed for
other chemical applications. Ultrasound has a greater influence on extraction at lower tem-
peratures where the energy generated upon cavitation bubble collapse is greater. Without
sufficient heat the solubility of the extracted materials may decrease.
UAE extends the range of solvent choice so that generally-regarded-as-safe (GRAS)
solvents may be used to replace organic solvents, which could reduce environmental impact
and cost as well as provide health and safety benefits.(10) One particular application is
the replacement of organic solvents with supercritical carbon dioxide in combination with
ultrasound.(15) Other demonstrated advantages of UAE include high reproducibility, high
level of automation, overall decrease in both solvent usage and consumption of fossil
Downloaded by [University of Windsor] at 02:26 17 June 2013

energy, elimination of posttreatment of waste water, and an increase in analyte purity.(16–18)


UAE is reported to be cost-effective, simple, and efficient. Moreover, it is highly attractive
from an industrial perspective, as it can be added onto an existing processes technology
with minimal alteration.
Despite its many benefits, however, optimization of the ultrasound operational param-
eters (sample size, particle size, solvent type and concentration, ultrasonic power and
frequency, extraction time and temperature, liquid/solid ratio) is crucial to achieve high
extraction efficiency, and the optimization conditions vary with the nature of the mate-
rial and the analyte of interest.(16,19) It is also important to appreciate that the conditions
that produce the greatest yield may be different to those that give rise to the extract with
maximal desired functionality.
The application of UAE has been used to isolate various classes of food components,
such as aromas, lipids and oils, pigments, polysaccharides and proteins,(20,21) metals,(22,23)
chemical contaminants,(24,25) foam stabilizers,(26,27) as well as a range of bioactive compo-
nents (e.g., anthocyanins, carotenoids, flavonoids, phenols). Several comprehensive reviews
in this area have been published.(8,10,16,17,28)

Ultrasound-Assisted Extraction of Polysaccharides


The benefit of using ultrasound to extract carbohydrates and polysaccharides of interest
to the food industry has been demonstrated for a range of substrates. Specific exam-
ples include the extraction of high-molecular-weight polysaccharides, such as chitin from
seafood waste,(29,30) cellulosic components from agricultural food waste,(31,32) and xyloglu-
can from apple pomace.(33) However, it is important to appreciate that sonochemical
modification (e.g., radical-induced fragmentation) of the polysaccharide may occur during
the process.(18) The modification may facilitate the extraction process (e.g., cleave linkages
between different biopolymers) or alter the properties of the biopolymer (e.g., reduce its
molecular weight) in either desirable or undesirable ways. For instance, the application of
ultrasound (intensity [I] = 41 W/cm2 , up to 4 h) during acid demineralization of ground
shrimp shell decreased the process time, enhanced deprotonation but not demineralization,
and resulted in significant extraction of chitin from the shells into solution due to fragmen-
tation of the biopolymer.(29) However, when ultrasonication was performed following alkali
demineralization, deprotonation was enhanced and chitin was transformed structurally for
more efficient subsequent processing.(29) In sugarcane bagasse (fibrous matter that remains
after sugar cane stalks are curshed to extract juice), UAE (P = 100 W, f = 20 kHz, 55 ◦ C,
70 Chandrapala et al.

40 min) expedited the extractability of hemicelluloses by apparent destruction of the plant


cell walls and cleavage of ether linkages between lignin and hemicellulose.(34) The decrease
in molecular weight of biopolymers is related to the structure (e.g., extent of branching) and
composition of the molecule, and has been attributed to the strong shear stresses that are
exerted during growth and collapse of cavitation bubbles.(18) Interestingly, aqueous UAE
(frequency not provided) extraction of polysaccharides from the edible fungus, Pleurotus
tuber regium, resulted in glycan-chitin complexes with a higher molecular weight than com-
pounds generated by hot water extraction,(35) possibly due to sonochemical modification
of the polysaccharides.(10) Furthermore, the polysaccharides obtained by UAE exhibited
enhanced antitumor and immunological activities in animal trials.(35)
Conventional solvent extraction is typically used to extract polysaccharides. Although
solvent extraction is simple, high temperature and long extraction time lead to decom-
position of polysaccharides and loss of their bioactivity. UAE (P = 60 W, water:solid
ratio of 15:1, 60 ◦ C, 20 min) extraction of bioactive polysaccharides from mulberry
Downloaded by [University of Windsor] at 02:26 17 June 2013

leaves was shown to more efficient than either solvent extraction or microwave-assisted
extraction, providing the largest yield of polysaccharides and requiring the least ratio of
water to raw material.(36) Importantly, the polysaccharides obtained by UAE displayed
similar physicochemical and structural characteristics to those obtained by the other
methods.
Assorted one- and two-step extraction procedures, with and without a short (typi-
cally 5–10 min) ultrasonic treatment applied at the start of the extraction, have been used
to enhance the extractability of polysaccharides from a range of food materials. These
include, for example, UAE extraction of polysaccharides from almond shells,(37) pectic
polysaccharides from the seeded fruit of oil pumpkin,(38) phenolic-rich heteroxylans from
wheat bran,(39) hemicelluloses from corn bran(31) and buckwheat hulls,(32) and starch from
cassava root pulp.(40) In all cases, either an increase in yield and/or shorter extraction time
was obtained by UAE compared with extraction in the absence of ultrasound (Table 1).

Ultrasound-Assisted Extraction of Phytochemicals


Plant-derived phytochemicals have been the focus of extensive research due to their health-
promoting bioactive properties. The application of UAE for extraction of phytochemicals
is well established. Ultrasonication can be used to extract both hydrophilic (e.g.,
anthocyanins, caffeine, tannins) and hydrophobic (e.g., β-carotene, lutein, lycopene)
phytochemicals, and from a wide range of agricultural materials.
Several reviews demonstrating the benefits of ultrasound for isolation of
phytochemicals of relevance to the food and nutraceutical industries have been
published.(8,10) More recent interest is on commercialization of UAE for bioactive extrac-
tion on an industrial scale. A specific example of successful commercialization of UAE is
in the wine industry, which employs a 32-kW unit to treat 50 m3 /h of must for extraction
of grape anthocyanins and color pigments during the fermentation process.(41) As another
example, Virot et al.(42) demonstrated the possible use of ultrasound (P = 150 W, f =
25 kHz) to extract polyphenols from apple pomace on an industrial scale. Using a 30-L
ultrasound extraction reactor, the authors showed that although the polyphenol yield was
equivalent to that obtained on the laboratory scale, it was 20% higher than the conven-
tional maceration procedure. Furthermore, UAE could use lower extraction temperatures,
shorten the extraction time, and enhance the final yield, providing considerably lower oper-
ating costs. The potential for UAE at pilot-scale was also realized. Other advantages of the
UAE process include the recycling of agricultural by-products such as rice, buckwheat and
Downloaded by [University of Windsor] at 02:26 17 June 2013

Table 1
Selected literature examples of ultrasound-assisted extraction (UAE) of food components in some food matrixes

Target Food matrix Approach Remarks Reference


Melanin Dried fruit bodies of UAE (40 kHz/250 W) Increased melanin yield to Zou et al.(21)
Auricularia auricula 85 mg/100 g
Atrazine and Simazine Honey UAE with benzene:water 92.3% and 94.2% recovery of Rezić et al.(24)
1:1 mixture compared with atrazine and simazine,
traditional shake-flask respectively, were achieved by
extraction method UAE in comparison with 72%
and 75.7% through the
traditional method.
Ginsenosides (saponins) Ginseng UAE (20 kHz/ 85 W) compared Extraction yield with UAE Wu et al.(26)
with Soxhlet extraction increased by 3-fold over
conventional extraction and

71
required only 2 h compared with
8 h.
Bioactives Quillaja saponaria UAE (20 kHz, horn type, Similar yield was achieved with Cares et al.(27)
(Quillay) 30–70 W) 20 min UAE at 20 ◦ C compared
with 3 h conventional extraction
at 60 ◦ C; lower extraction
temperature, shorter time, and
increased yields were achieved
by UAE.
Noncellulose (xylan Corn bran Two-step extraction 10–40% increase in yield; xylan Ebringerová and
component/polysaccharide with/without horn-type (neutral sugar component) yield Hromádková(31)
fraction) ultrasound (100 W, 8 W/cm2 varied from 65% to 88% with
intensity) the use of sonication.

(Continued)
Downloaded by [University of Windsor] at 02:26 17 June 2013

Table 1
(Continued)

Target Food matrix Approach Remarks Reference


Hemicellulose Bucket wheat hulls Two-step extraction The yield of 3% NaOH-isolated Hromádková
with/without horn-type hemicelluloses increased by a et al.(32)
ultrasound (100 W, 8 W/cm2 factor of 1.4 and that of 5%
intensity) NaOH by 1.2 when compared
with standard extraction.
Pectin polysaccharides Oil pumpkin Two-step extraction Extraction time was reduced from Košt’álová et al.(38)
(Cucurbitapepo L. with/without horn-type 60 min to 5–10 min with the use

72
var. styriaca) ultrasound (100 W, 8 W/cm2 of UAE.
intensity)
Hemicellulose Wheat bran Two-step extraction Similar yields of polysaccharides Hromádková
with/without horn-type were achieved using 10 min et al.(39)
ultrasound (100 W, 8 W/cm2 UAE compared with 60 min
intensity) with conventional method.
Starch Cassava Use of ultrasonic power model Improved recovery efficiency by Sriroth et al.(40)
SL at 80% max power 40% by disrupting the complex
structure of polysaccharides
associated with the starch
granules.
Use of Ultrasound in Food Processing 73

almond hulls, straw and fruit peels, for example, and the fact that no additional treatment
and/or chemicals are required.
UAE has been employed to extract antioxidants from rosemary.(43) The longer the
extraction time (15–60 min) at 40 kHz and 50 ◦ C, the greater the extraction yield of carnosic
and rosmarinic acids. A comparison was made between extraction in a 20-L (5 g solid) and
a 125-L (6.25 kg solid) ultrasonic cleaning bath (f = 40 kHz, 35 ◦ C, up to 60 min) with ther-
mal extraction at 35 ◦ C, for up to 60 min. The 125-L ultrasonic bath allowed for 66% more
rosmarinic acid to be extracted (mg/g) when compared with the thermal process at 35 ◦ C
for 60 min (Fig. 1). The shorter the sonication extraction time, the greater the efficiency
of extraction, with 80% and 100% greater yields observed at 45 and 30 min, respectively,
when compared with thermal values. It was shown that the solvent employed was important
when tailoring ultrasonic extraction processes to a specific material. By adding water to the
system, the antioxidant levels were reduced due to poor solubility and degradation by OH•
and COO• radical formation.
Downloaded by [University of Windsor] at 02:26 17 June 2013

UAE can have either positive or negative effects on the quantity and antioxidant activ-
ity of bioactives. This is due to the potential for cavitation to generate free radicals, notably
OH• radicals, which may trigger or accelerate chemical reactions.(44) Ashokkumar et al.(44)
have shown the potential of high-frequency ultrasound (f = 358 and 1062 kHz) to enhance
the antioxidant activity of phenolic compounds. Sonication of phenol in aqueous solutions

Figure 1. Yield of carnosic (CA) and rosmarinic acid (RA) from rosemary leaves observed following
extraction at 35 ◦ C as a function of extraction time. Th = thermal bath extraction, no ultrasound.
Ultrasound-assisted extraction scaled up to 20 L volume (containing 1 kg dried rosemary) and 125 L
(containing 6.25 kg dried rosemary) volume at frequency of 40 kHz (from Paniwnyk et al.(43) ) (color
figure available online).
74 Chandrapala et al.

at different ultrasonic frequencies (I = 0.9 W/cm2 , up to 60 min) showed no significant


change in the concentration of phenol (1 mM) at 20 kHz, whereas there was a significant
decrease in phenol concentration at 358 and 1062 kHz. Analysis of the sonicated solutions
by high-performance liquid chromataography showed that a decrease in phenol concentra-
tion correlated with an increase in hydroxylated products. Also, a decrease in OH• radical
yield observed at 358 and 1062 kHz closely correlated with the decrease in phenol degra-
dation. It is known that an increase in the ultrasound frequency increases the number of
active cavitation bubbles.(6) Furthermore, at 20 kHz, the cavitation bubbles are principally
transient, whereas stable bubbles (repetitive transient cavitation bubbles)(6) are generated
at higher frequencies. Both an increase in the number of active bubbles and repetitive
transient cavitation can be anticipated to increase the amount of OH• radicals generated
with an increase in ultrasonic frequency. However, other studies have shown that prolonged
ultrasonic treatment times and high temperatures can have a significant adverse effect on
the antioxidant activity of phenolic compounds.(45,46) Tiwari et al.(47) observed degradation
of anthocyanins in grape juice during sonication (f = 20 kHz, pulsed 5 s on and 5 s off,
Downloaded by [University of Windsor] at 02:26 17 June 2013

25 ◦ C, 0–10 min) at amplitude levels >24.4 μm and this was attributed to sonochemical
and sonomechanical effects.
Where free radicals can adversely affect the quality of the extract, techniques have been
developed to control free radical reactions. For example, to avoid or minimize oxidative
damage, radical production can be quenched by addition of a radical quencher, such as
ascorbic acid.(44) Alternatively, degassing has been used to enhance the antioxidant activity
of an extract obtained from sonication of bearberry leaves.(45)

Ultrasound-Assisted Extraction of Oils


Ultrasound is emerging as a valuable processing aid in the extraction of oil. Wei et al.(48)
compared extraction of oil from rapeseed by ultrasound (P = 500 W, f = 40 kHz, maximum
35 ◦ C, 60 min) with standard solvent extraction and found that the use of ultrasound reduced
the extraction time 5-fold. Furthermore, both the total oil content and fatty acids in the UAE
extract agreed with that obtained by the standard method.
Lou et al.(49) described a dynamic UAE procedure (P = 250 W, f = 40 kHz, 5 mL/min
flow rate, solvent:solid ratio of 8.5, 50 ◦ C, 20 min) that resulted in a respective 2% and 10%
higher oil yield from chickpea than those obtained by static UAE and solvent extraction.
In addition, dynamic UAE was a respective 12- and 15-fold faster than solvent extraction
and static UAE, and used less solvent and a lower extraction temperature. Extraction at
higher temperatures results in a substantial loss of tea aroma and causes undesirable alter-
ation in the characteristic flavor and appearance of tea. Extraction at lower temperatures
would therefore be preferable but this suffers from the disadvantage that it results in low
extraction efficiency. It was also shown that dynamic UAE did not significantly alter the
composition of the oil. Similar findings were reported by Lin et al.(50) comparing UAE
with solvent extraction of oil from hempseed. UAE (P = 200 W, f = 20 kHz, pulsed 20 s
on and 20 s off, solvent:solid ratio of 7, 30 min) decreased the extraction time and con-
sumed less solvent while not significantly altering the oil composition. Interestingly, the
UAE oil extract also displayed 13.5% higher radical scavenging activity in vitro than that
obtained with solvent extraction. Chemat et al.(51) used 20 kHz (P = 150 W, 69 ◦ C) UAE to
extract carvone and limonene from caraway seeds and compared extraction efficiency with
Soxhlet and conventional solid-liquid extraction. Extraction rates were increased 1.3–2-
fold by UAE compared with the conventional methods. The UAE extracts were also of
higher quality and purity than either of the other extraction methods. Scanning electron
Use of Ultrasound in Food Processing 75

microscopy (SEM) images confirmed the efficiency of UAE was due to sonication-induced
mechanical disruption to the plant cells. Although the studies described for UAE oil extrac-
tion have principally been performed on plant material, it may also be applied to material
from animals. A recent study by Abdullah et al.(52) showed the potential of UAE (P =
200 W, f = 25 kHz, solvent:solid ratio of 50, 60 min) to extract oil (7%) from eel.

Ultrasound-Assisted Extraction of Proteins


Rennin is an industrially important food enzyme that plays a key role in milk coagula-
tion for cheese production. Zayas(53) reported that sonication (I = 3.34 W/cm2 , 15 ◦ C, 5%
NaCl, solid:liquid ratio of 1:18, 45 min) increased the yield and activity of rennin from abo-
masa of calves in comparison with a standard extraction method.(53) More recently, Zhang
and Wang(54) optimized the ultrasonic processing parameters (I = 44 W/cm3 sonication
bath, 16% NaCl, solid:liquid ratio of 1:30, 25 min) for extraction of goat kid rennet, and
Downloaded by [University of Windsor] at 02:26 17 June 2013

recommended conditions different from those of Zayas,(53) possibly due to differences in


tissue structure between animal species. The UAE process increased extraction efficiency
by ∼43% compared with extraction and significantly shortened the extraction time.(54)
Ultrasonic extraction has been used to recover valuable protein from several by-
products of the food industry. Brewer’s spent grain is a high protein-containing by-product
of the brewery industry. Tang et al.(55) optimized the UAE conditions (P = 88.2 W/100 mL,
solid:liquid ratio of 2:100, ∼20 ◦ C, 81.4 min) for extraction of protein from brewer’s grain.
The experimental yield was in agreement with the predicted value.(55) Zhang et al.(56) estab-
lished an ultrasonic-assisted (P = 160 W, f = 40 kHz) method to isolate protein from ground
almond dregs. Extraction time was found to be a major determinant of the amount of pro-
tein extraction. The optimal efficiency of protein extraction was 70% at 37 ◦ C with 2 ×
30 min exposure times (pH 10.1, solid:liquid ratio of 1:29.5).
It is noteworthy that ultrasound can induce changes to protein structure. Amino acids
having sulfhydryl and phenolic residues can be modified by OH• radicals, generated by
cavitation bubbles, to form new covalent linkages between proteins.(57) Also, multimeric
globular proteins may be broken down into subunits, and if the exposure to ultrasound is
too severe, protein hydrolysis may occur.(57) Such structural changes can affect the protein
functional properties. Karki et al.(58) showed that pretreatment of defatted soy protein flakes
with ultrasound at low power (f = 20 kHz, I = 0.30 W/mL, 30–120 s) and high power (f =
20 kHz, I = 2.56 W/mL, 30–120 s) gave increased yields of soy protein isolate of 13%
and 34%, respectively. UAE at high power (f = 20 kHz, I = 2.56 W/mL, 120 s, 65 ◦ C)
improved soy protein isolate solubility by 34% at pH 7 and decreased its emulsification
(by 12%) and foaming (by 26%) capabilities compared with control soy protein isolate.
Rheological behavior of the protein was also modified with a significant decrease in con-
sistency coefficient. These changes were attributed to protein denaturation with increased
sonication time and power. Similarly, Chen et al.(59) found that the extraction yield of soy-
bean protein isolated from soybean flakes was improved with ultrasound (P = 300 W,
f = 12 kHz, 8 min). Protein extraction yields obtained by ultrasonic treatment of defat-
ted soy flakes and white flakes were respectively ∼73% and ∼90%, compared with ∼50%
and ∼67% by conventional stirring extraction at 45 ◦ C for 1 h. The ultrasound extrac-
tion procedure simultaneously altered the protein properties. In particular, the emulsifying
property of soy protein isolate increased whereas the viscosity decreased with ultrasonic
treatment. In a previous study, Wang(60) found that ultrasonic treatment (P = 125 W, f =
20 kHz, 1:10 solid:liquid, 8 min) of defatted soybean flakes tended to promote soy pro-
tein aggregation rather than dissociation in water extracts. The mechanism of aggregate
76 Chandrapala et al.

formation by sonication was not revealed. However, it is possible that ultrasound may pro-
mote hydrophobic interactions of globular proteins in water, may modify the equilibrium
condition of protein-lipid or protein-protein interactions to facilitate cluster formation, or
may induce complex formation, as in apolipoproteins.(61)

Ultrasound-Assisted Extraction of Aroma and Flavor Volatiles


Aromas and flavors are complex mixtures of volatile compounds typically present at low
concentrations and exist in various aromatic plant materials. Conventional flavor/aroma
extraction techniques have major drawbacks, such as low yields and the formation of
by-products, due to the low stability of the target compounds. In contrast, the mild pro-
cessing conditions (temperature and extraction time) of ultrasound can lead to enhanced
extraction efficiency of sensitive food components. Jadhav et al.(62) compared the extrac-
tion efficiency of vanillin from vanilla pods by UAE (P = 240 W, f = 22.4 kHz, 5 s off
Downloaded by [University of Windsor] at 02:26 17 June 2013

and 5 s on, solid:solvent 1:100) with Soxhlet extraction. Under optimal Soxhlet condi-
tions, about 180 ppm of vanillin was released in 8 h, whereas UAE required 1 h to extract
140 ppm vanillin at similar solvent:solute ratio and at ambient temperature. UAE (typically
f = 40–48 kHz) has enabled rapid quantification of various volatiles in beer, wine, and
other alcoholic beverages.(63,64)
Alissandrakis et al.(65) compared UAE (frequency not provided, 25 ◦ C, 10 min)
with various extraction techniques (hydrodistillation, solid-phase microextraction, microsi-
multaneous steam distillation–solvent extraction, and high-hydrostatic-pressure treatment)
for honey aroma compounds. Wide variability in the aroma compounds was observed
among the different techniques employed. Hydrodistillation and microsimultaneous steam
distillation–solvent extraction destroyed sensitive compounds and led to the formation of
artifacts. In contrast, UAE at ambient temperature avoided any thermal degradation and
resulted in an increased yield of aroma volatiles. Moreover, both low- and high-molecular-
weight compounds could be extracted by UAE, providing a useful fingerprint for honey
origin control. UAE (f = 35 kHz, 25 ◦ C) also caused less damage to thermosensitive aroma
compounds in garlic compared with microwave-assisted extraction-hydrodistillation and
standard extraction, due to the mild sonication conditions.(66)
Extraction is one of the essential processes in the preparation of key food ingredients.
It plays a vital role in optimizing the quality of common foods and beverages. There is also
increasing interest in the use of extraction for the production of high-value compounds,
such as bioactive peptides from milk and whey, and the recovery of co-products from food
processing wastes. Unlike some other nonthermal processes, such as high-pressure process-
ing, pulsed electric field, and either compressed or supercritical carbon dioxide extraction,
ultrasound can be easily tested in laboratory or bench-top scale, generating reproducible
results for scale-up. The intensity and the cavitation characteristics can be easily adapted
to the specific extraction process to target specific objectives. Amplitude and pressure can
be varied over a wide range, for example, to identify the most energy efficient extraction
setup. Ultrasound can also act synergistically with temperature and pressure to achieve
higher benefits.

Crystallization
The crystallization of biological soft materials with the use of ultrasound plays a key role
in controlling the crystal structure and shape and rate of crystallization. Sonocrystallization
is the use of power ultrasound to control the crystallization process commonly used during
Use of Ultrasound in Food Processing 77

the nucleation phase of crystallization. Rapid sonocrystallization for lactose recovery, fat,
and proteins in the food industries relies on different aspects of ultrasound.
The main problem associated with lactose crystallization is that it has a long induction
time and slow crystallization rate due to a large metastable zone width. Optimization of
process parameters for rapid recovery of lactose by ultrasound-assisted crystallization in
the food industry has been investigated.(67–69) A dramatic shortening of the induction times
(an increase in the corresponding nucleation rates) observed in the sonicated systems has
been attributed to two major factors: ultrasound-stimulated nucleation in the high-pressure
vicinity of the collapsing bubbles due to the melting-point shift and the consequent increase
in supercooling.(67) Ultrasound in solution generates cavitation bubbles, which have been
associated with instantaneous high pressures and temperatures. These cavitation spots are
thought to be privileged nucleation centers, where the critical energy for crystal formation
is decreased. At the same time, the acoustic waves may increase both the probability of
contact between molecules and mass transfer, thus increasing the nucleation frequency.
Downloaded by [University of Windsor] at 02:26 17 June 2013

Hence, ultrasound is a promising tool for rapid crystallization of food materials, which
will increase the efficiency of some traditional processes, leading to cost effectiveness.(68)
Ultrasound not only reduces the induction time at low super saturations, but also narrows
the metastable zone width. Ultrasound can be used to induce nucleation where spontaneous
primary nucleation cannot occur and hence, ultrasound can potentially replace seeding in
solution.
In the study of Bund and Pandit,(67,68) the whole process of lactose sonocrystallization
was rapidly completed from reconstituted lactose solutions in the presence of “ethanol” as
an anti-solvent, at 30 ± 2 ◦ C. The lactose recovery was much higher (91%) in the case of
sonicated samples when compared with nonsonicated (14%) samples at the end of 5 min.
As ultrasound travels throughout the crystallization vessel, it mixes the anti-solvent uni-
formly in the lactose solution. Cavitation events are also likely to be increased due to the
lowering of the vapor pressure of the solution as a result of the addition of ethanol. This
causes a sharp decrease in the local solubility of lactose. The solution reaches supersatu-
ration, causing nucleation and rapid lactose recovery. The effect of cavitation to speed up
the nucleation rate is a complex phenomenon and is not completely understood. It could
be hypothesized that every cavity takes up some part of the solvent, causing instantaneous
local supersaturation. It could also be speculated that due to hot spot formation, as a result
of cavity collapse, local solvent evaporation takes place within the cavitation nuclei. On col-
lapse of a cavity, these nuclei are introduced into the crystallization slurry, causing further
cavitation opportunities, due to the presence of these solids.(70)
A further benefit of ultrasound is the improvement in crystal quality as a result of
reducing agglomeration and changing crystal habit.(67–69) The lactose crystals recovered
from reconstituted lactose solution (lactose content 17.5% w/v, protein content 0%), in the
presence of sonication (Fig. 2a) were small and uniform in shape and showed no aggre-
gation compared with those obtained in the absence of sonication (stirring) (Fig. 2b) at
the end of 5 min of crystallization time. Some lactose crystals recovered without sonication
showed broken corners and dendritic forms of crystals (Fig. 2b). Hence, sonocrystallization
not only enables rapid crystallization but also guarantees a relative uniformity of the crys-
tal size distribution and prevention of agglomeration in comparison with nonsonicated
samples. It was observed that the maximum percentage of crystals were in the range of
2.5–6.5 μm at the optimized conditions, whereas the pure lactose crystals were in the range
of 3.5–9.5 μm.(67) Patel and Murthy(69) also investigated the shape characteristics and size
distribution of lactose crystals recovered with the use of ultrasound and observed similar
findings.
78 Chandrapala et al.

Figure 2. Lactose (L) crystals observed under the microscope (40× objective): (a) sonicated and (b)
nonsonicated lactose solutions (17.5% w/v) (from Bund and Pandit(67) ).
Downloaded by [University of Windsor] at 02:26 17 June 2013

There has been a considerable interest in the application of ultrasound with regards
to crystallization of fats. Examinations of pure triacylglycerides (TAGs), confectionary
fats, vegetable fats, and milk fats indicate that ultrasound affects the rate of polymorph-
dependent crystallization, as well as crystal size and morphology.(71–76) The crystallization
behavior of tripalmitin (PPP) and trilaurin (LLL), without and with the application of
ultrasonic power, was investigated by Ueno et al.(77) The following effects were observed
from their study at initial crystallization temperatures of 50 and 30 ◦ C and ultrasound
applied for 2 s: (a) a marked decrease in induction times for crystallization of both PPP
and LLL; (b) an increased nucleation rate; and (c) crystallization of only the β form for
both PPP and LLL. The last finding demonstrated that ultrasound irradiation can be used
as an efficient tool to control the polymorphic crystallization of fats. In addition to this, the
crystallization of LLL under a lower initial crystallization temperature of 25 ◦ C and with
the same ultrasonication time of 2 s revealed the presence of both α and β polymorphs.
The authors further suggested that crystallization of only the β form depends not only on
ultrasound treatment but also on the initial temperature of crystallization. On the basis of
the dynamic nucleation of PPP and LLL crystals induced by collapsing cavitation bub-
bles, the authors argued that the pronounced decline in induction times and increase in the
nucleation rate resulted from a shift in the melting points of the fats due to the high-pressure
pulses associated with collapsing bubbles. More importantly, this procedure may suggest
a rational way of controlling the polymorphism and nucleation rate of various fats/lipids
by sonication in food-based systems. Ye et al.(76) used high-intensity ultrasound (HIU) to
change the crystallization behavior, generate small crystals, and improve the texture of
intersterified soybean oil. Samples were crystallized at different temperatures (26–32 ◦ C)
with/without the application of HIU at different power levels (P = 110, 72, 61, 54, and
44 W). Their results showed that higher acoustic power had a greater effect on crystal size
reduction, induced crystallization, and generated harder, more elastic and viscous materi-
als. HIU induces primary and secondary nucleation and not only promotes crystal growth,
but also has the function of stabilizing the crystals that are formed. The effectiveness of
this technology depends entirely on the processing conditions. The studies reviewed above
indicate that ultrasound affects the crystallization behavior of fats in many ways. However,
it is still unclear what exact mechanisms are responsible for these effects. To improve
our knowledge and predictability in terms of desired polymorphism, induction times, and
nucleation rates as influenced by sonocrystallization, a better understanding of the phase
diagram for polymorphic forms is required. Indeed, a primary effect of sonocrystallization
may be due to (i) the high pressure generated when a sonication-induced cavity is collapsed;
Use of Ultrasound in Food Processing 79

(ii) stability (lifetime) of different polymorphic forms as a function of supercooling tem-


perature; (iii) mechanism and lifetime of collapsing cavities; and (iv) the basic mechanism
of the dynamic nucleation, in the vicinity of a collapsing bubble.(74)
Producing high-quality crystals has often been an obstacle to protein structure analy-
sis. Various means of inducing nucleation have been investigated by many scientists.(78–80)
The ability of ultrasound to reduce the metastable zone width in pharmaceutical crystal-
lization processes has been adapted for use as a protein crystallization promoter in the food
industry. Crespo et al.(81) studied the crystallization of egg-white lysozyme in the pres-
ence and absence of ultrasonic irradiation using commercial crystallization plates placed in
temperature-controlled water baths. By inducing faster nucleation, ultrasound led to protein
crystals grown at low supersaturation levels; such crystals are known to have better diffrac-
tion properties. By increasing the nucleation probability, the presence of ultrasound shifted
the metastable curve to regions of the phase diagram of lower supersaturation levels and
therefore the metastable zone width decreased. As a result, crystallization conditions that
Downloaded by [University of Windsor] at 02:26 17 June 2013

do not normally produce crystals may become successful in the presence of ultrasound.
This can be particularly advantageous for proteins showing a very limited range of con-
ditions corresponding to the nucleation zone. These results confirmed the suggestion that
ultrasonic irradiation can be advantageously used in protein crystallization, not only as a
nucleation promoter, which decreases the energy barrier for crystal formation, but also to
enhance the quality of protein crystals.

Thawing
The process of thawing bulky frozen food is extremely slow. Hence, researchers have
focused on alternative ways to generate heat within the food to accelerate the thawing
process. Microwave, dielectric, or resistive heating methods have severe restrictions due
to run-away heating and preferential surface heating. If the temperature of the product is
nonuniform, warm regions become centers for more power dissipation. When this happens,
run-away heating may result in regions of cooked product embedded within substantially
frozen product. Preferential heating of regions of the product near the surface arise due
to the higher intensity of the radiation at the surface and falls exponentially with depth
due to absorption. Surface regions therefore tend to thaw first.(82) Hence, ultrasound has
been considered as an alternative thawing method. It has been shown that ultrasound is
more highly attenuated in frozen meat than in unfrozen tissue,(83) and that the attenuation
increases markedly with temperature, reaching a maximum near the initial freezing point of
the food.(82) Thus, in contrast to microwave, the ultrasound attenuation-temperature profile
is well suited to producing uniform and rapid thawing. The product was irradiated from the
outside in order to control the energy input and hence the rate of thawing.(82)
Miles et al.(82) undertook ultrasonic thawing tests on frozen meat and fish samples to
assess the effectiveness of HIU for thawing of frozen foods. Previous attempts by other
investigators(84,85) used ultrasound merely to assist the thawing of material immersed in
warm water. In the approach of Miles et al.,(82) ultrasound supplied all the energy require-
ments, whereas in other systems(84,85) almost all the thawing energy came from the warm
water in which the material was immersed, and sonic and ultrasonic vibration merely
assisted heat transfer into the material. The times taken to thaw to specified depths, at
frequencies of 415, 460, 470, 520, and 740 kHz and intensities of 0.5–0.7 W/cm2 for
various meat and fish samples were compared.(82) The observed thawing times were con-
siderably faster than that obtained by thermal conduction from a surface held at a constant
temperature of 25 ◦ C.
80 Chandrapala et al.

As the ultrasonic wave is transmitted through a slab of frozen meat, it loses energy
and much of this is dissipated as heat, causing the temperature to rise rapidly at the sur-
face and less rapidly internally. As the temperature rises, the attenuation and the rate of
energy dissipation per unit volume also increase. However, the specific heat capacity, and
consequently the rate of temperature rise, may not vary markedly. As the initial freezing
point is approached, more and more energy will be dissipated and little energy will pass
into the bulk of the meat. Eventually the initial freezing point will be exceeded, and the
attenuation will fall dramatically at the surface but will be maintained at the very high level
within a region of the frozen tissue close to the thawed/frozen boundary. This boundary
can be pushed into the tissue with control at a rate dependent on the intensity of ultrasound
applied. Miles et al.(82) found that using either low frequencies (f < 430 kHz) at intensities
of 0.5–2 W/cm2 , or frequencies ≥740 kHz, overheating and poor ultrasonic penetration
was achieved. However, at frequencies around 500 kHz and intensity around 0.5 W/cm2 ,
acceptable surface temperatures were obtained. Furthermore, a theoretical model indicated
Downloaded by [University of Windsor] at 02:26 17 June 2013

that thawing times could be reduced by suppressing cavitation and using lower frequencies
at higher intensities. Use of ultrasound in thawing frozen foods is a noninvasive and rapid
approach, which is beneficial for the food processing industry.

Drying
Drying is commonly used to preserve foods and is frequently based on the use of thermal
energy.(86–88) Conventional air-drying is energy- and cost-intensive because it is a simul-
taneous heat and mass transfer process accompanied by a phase change (liquid to vapor).
Moreover, drying generally produces structural changes in the products and can deteriorate
the quality of the final product, such as forming undesired food flavors and colors, vitamin
degradation, and the loss of essential amino acids.(86) For fruits, a pretreatment process,
such as osmotic dehydration, is often used to either reduce the initial water content of the
fruit or to modify the fruit tissue structure in such a way that air-drying becomes faster.(87)
A great emphasis is presently given to novel treatments where the quality will be preserved
and enable the introduction of new, safer, fresher, and better-quality foods with longer shelf
life. Among emerging new technologies, ultrasonic dehydration is very promising because
the effects of power ultrasound are more significant at low temperature, which reduces
the probability of food degradation. In addition, ultrasound permits the removal of mois-
ture from solids without producing a liquid-vapor phase change. Hence, the application of
ultrasound in drying processes, which represents an emergent and promising technology,
has been investigated by many researchers.(86–90)
The application of ultrasound could constitute a method of improving traditional con-
vective drying systems due to the following. When a high-intensity ultrasonic wave is
directly coupled to the material to be dried, it travels through the solid medium, causing a
rapid series of alternative compressions and expansions, in a similar way to a sponge when
it is squeezed and released repeatedly (known as the sponge effect). The forces involved
in this mechanical mechanism can be higher than the surface tension, which maintains
the moisture inside the capillaries of the material, creating microscopic channels that may
facilitate moisture removal. Other effects to be considered are the variation of viscosity and
surface tension and deformation of the porous solid material.(86–91)
As an example, Fernandes et al.(87) evaluated the effect of osmotic dehydration and
ultrasound pretreatments on melon tissue. The influence of both pretreatments on water
diffusivity in the air-drying step was analyzed. The changes in the value of the water diffu-
sivity had great significance during the air-drying stage for both treatments. For example,
Use of Ultrasound in Food Processing 81

if melons were dried from an initial moisture content of 9.10 g of water/g of solids to a
final moisture content of 0.25 g of water/g of solids, it would require 760 min to dry the
melons, based on a water diffusivity of 5 nm2 /s. Submitting the melons to 30 min of ultra-
sound treatment reduced the drying time to 550 min due to an increase in water diffusivity
to 7 nm2 /s. Submitting the melons to 1 h of osmotic dehydration, the drying time would
be reduced to 605 min due to the increase in water diffusivity to 6.3 nm2 /s. However, both
ultrasound and osmotic drying treatments induced changes to melon cell structure, although
no cell breakdown was observed with the former. The effective water diffusivity in fruits is
dependent on tissue structure, since the cell walls act as a semi-permeable membrane, and
also on the fruit porosity. Large intercellular spaces can be found in high-porosity products,
making the product more prone to alternating compression and expansion cycles produced
by ultrasound waves (sponge effect) and facilitating the water transfer through the solid.
The effect of porosity could also be explained by considering the acoustic energy absorp-
tion in high-porosity products due to larger gas volume, which may produce an increase in
Downloaded by [University of Windsor] at 02:26 17 June 2013

the internal energy available in the particles. Moreover, the intercellular spaces increase in
high-porosity products due to larger pores in the food matrix, and this in turn could reduce
the internal mass transfer. The pretreatments studied by Fernandes et al.(87) increased the
effective water diffusivity of melons through different effects. Ultrasonic waves created
microscopic channels in the fruit, which increased the effective water diffusivity because
water could use these microscopic channels as an easier pathway to diffuse towards the
surface of the fruit.
The microscopic image analysis of fresh fruit showed typical thin-walled cells with
normal morphology and some intercellular spaces (Fig. 3A).(87) Very few cells had undu-
lated walls. Microscopic image analysis of 20-min-sonicated samples showed two distinct
regions, one region where the cells became bloated as a result of water gain (Fig. 3B)
and a second region where the cells became much smaller and needle-shaped (Fig. 3C).
These needle-shaped cells were not observed in the fresh fruit or in the samples submitted
to osmotic dehydration but were observed in all samples submitted to ultrasound pretreat-
ment. As shown in Fig. 3C, samples that were pretreated by ultrasound contained regions
of microscopic channels formed by the elongation and flattening of the cells. This result
confirmed the observations of Fuente-Blanco et al.,(86) which showed that the ultrasonic
pretreatment affects the fruit tissue and facilitates water diffusion during air-drying. The
generation of microscopic channels in the plant tissue may contribute to the increased
water diffusivity. After 30 min of ultrasonic treatment, the microscopic channels became
broader (Fig. 3D) and some new smaller microscopic channels appeared. A region contain-
ing bloated cells was also observed (Fig. 3E). The broadening of the microscopic channels
may further explain the increase in water diffusivity of the fruit in the air-drying process.
There was no indication of loss of the cell wall strength, and during knife cutting the cells at
the border did not fail under stress. Furthermore, in contrast to that observed when the fruit
was submitted to osmotic dehydration, no cell disintegration was observed in the samples
submitted to ultrasound.
Fernandes et al.(89) observed similar water diffusivity behavior on drying of
ultrasonically pretreated melon tissues to that of pretreated pineapple tissues. However, the
sonication-induced formation of microscopic channels in the pineapple tissues was primar-
ily attributed to loss of cellular adhesion (disruption of contiguous cells), which produced
large cell interspaces. Hence, the formation mechanism of microscopic channels in pineap-
ples differed from the mechanism observed in melons,(87) where microscopic channels were
formed by flattening and elongation of cells. Azoubel et al.(88) found that it took around
345 min to dry bananas at 50 ◦ C to a final moisture level of 0.33 kg H2 O/kg dry matter
82 Chandrapala et al.

B D
Downloaded by [University of Windsor] at 02:26 17 June 2013

C E

Figure 3. Photomicrographs of melon cubes before processing (A); after 20 min of ultrasound pre-
treatment, showing region with bloated cells (B) and region with microscopic channels (C); and after
30 min of ultrasound pretreatment, showing region with bloated cells (D) and region with microscopic
channels (E). Ultrasonic pretreatment parameters: 25 kHz frequency, 4270 W/m2 acoustic intensity,
4:1 water:fruit weight ratio, 30 ◦ C (from Fernandes et al.(87) ).

(25%, wet basis), which is the maximum moisture value permitted in dried fruits. However,
by subjecting the fruit to a 20-min ultrasound pretreatment, in distilled water, the drying
time was reduced to 207 min, as a result of an increase in water diffusivity.
Garcia-Perez et al.(91) investigated the influence of ultrasound on the convective drying
of carrot and lemon peel. In both products, the application of ultrasound improved the effec-
tive water diffusivity. The improvement was linearly proportional to the applied acoustic
power density. In the case of carrots, the average effective moisture diffusivity values identi-
fied ranged from 1.2 × 10−10 (I = 8 kW/m3 ) to 1.7 × 10−10 (I = 37 kW/m3 ) m2 /s, whereas
in the case of lemon peel, these values varied from 5 × 10−10 (I = 4 kW/m3 ) to 11.5 ×
10−10 (I = 37 kW/m3 ) m2 /s. The alternating expansions and compressions produced by
power ultrasound in the material being dried would be more intense as more acoustic energy
was applied, thereby improving the water transfer to the surface through the solid. Thus, the
improvement in effective moisture diffusivity by ultrasonic application could be explained
Use of Ultrasound in Food Processing 83

by considering the aforementioned influence on internal mass transfer.(91) The microstream-


ing, pressure variations, and oscillating velocities produced by the ultrasound on interfaces,
leading to a reduction in boundary layer thickness, are responsible for the effects on the
external resistance. The fact that lemon peel and carrots behave in a different way when
acoustic energy is applied could be linked to differences in their cell structure. Indeed,
lemon peel is considered to be a more porous product than carrot. Therefore, porosity may
be considered as one of the most important structural variables for determining the acoustic
effectiveness in foods. Large intercellular spaces are found in high-porosity products, mak-
ing the product more prone to alternating compression and expansion cycles produced by
ultrasonic waves (sponge effect) and facilitating the water transfer through the solid. The
small intercellular spaces, characteristic of low-porosity products, produce higher resis-
tance to internal water transfer. As a consequence, high acoustic energy levels are required
to affect mass transfer in carrot, which is considered as a low-porosity product. Greater
acoustic energy absorption could also be expected in high-porosity products due to a larger
Downloaded by [University of Windsor] at 02:26 17 June 2013

gas volume, producing an increase in the internal energy available in the particles. This
fact would lead to more intense compressions and expansions. Furthermore, the acoustic
effects on the boundary layer of intercellular spaces could be more intense in high-porosity
products due to a larger porous cell network. Indeed, these phenomena may also contribute
to reducing internal resistance to mass transfer.(91)
As the studies cited in this section illustrate, ultrasound is a promising emerging tech-
nology that may be applied to drying in order to reduce the processing time and thereby
represents an economy of energy in industry applications.

Freezing
Freezing is an alternative method for preservation of food. Preservation through freez-
ing can provide food with improved taste, texture, and nutritional value relative to other
preservation methods. Hence, increasing the production efficiency and quality of foods by
freezing has attracted much attention.(92) The presence of large ice crystals within the food
results is a major drawback in using freezing as a preservation technique, as it reduces the
eating quality of the product, including appearance, sensory properties, textual attributes,
and nutritional value. The freezing rate governs the formation of ice crystals, which in turn
dictates the preservation of the cellular structure. Faster freezing rate produces numerous
small ice crystals evenly distributed throughout the product, resulting in less damage to
cellular structure. On the contrary, a slow freezing rate leads to few but large ice crystals,
which causes breakage of the cellular structure.(93) Hence, the use of ultrasound to hasten
the freezing rate has been explored by some researchers.(92–94)
During ultrasonication, samples are exposed to an increasing amount of heat as a func-
tion of an increase in sonication time, due to the thermal effect of ultrasound. If the cooling
unit is not capable of removing the heat immediately, the amount of heat generated under
high acoustic power levels is large enough to raise the temperature of the refrigerating
medium. As a result, the heat transfer in the freezing process would be retarded. In this
case, the thermal effect would undermine the advantage of ultrasonic enhancement in heat
transfer during the freezing process. Hence, it has been suggested that the ultrasonic power
level applied (P = 0–300 W) should be chosen under the consideration of combining the
thermal effect of ultrasound and its enhancing effect on heat transfer.(92) As an example,
Li and Sun(92) explored the application of power ultrasound (f = 25 kHz) in immersion
freezing of potato. The objectives of their study were to evaluate factors influencing the
effects of power ultrasound on freezing efficiency in immersion freezing using optimized
84 Chandrapala et al.

operation conditions. It was found that during ultrasound irradiation (in the temperature
range of 0–5 ◦ C and power of 15.85 W, the temperature of the potato samples dropped
fastest for an exposure time of 2.5 min. At longer ultrasound exposure time (>2.5 min),
the freezing process was hindered and the samples required a longer time to reach the final
temperature. Furthermore, the effects of power ultrasound on the three distinct phases of
the freezing process (chilling, phase change, and tempering), was investigated by Li and
Sun.(92) Interestingly, when power ultrasound (P = 15.85 W) was applied to the tempering
phase (tef = 7.6 min), the freezing rate was almost equal to that without ultrasound (tef =
7.7 min). When ultrasound was present during the chilling phase, the freezing rate (tef =
7.2 min) was improved, but was not significantly different from that without ultrasound.
When power ultrasound (P = 15.85 W) was applied to the phase-change period, the freez-
ing process was significantly improved and the fastest freezing rate (tef = 6.6 min) was
obtained. During the phase-change period, a large amount of latent heat (335 kJ/kg) gener-
ated from the transformation of water into ice should be removed.(95) Most of the freezable
Downloaded by [University of Windsor] at 02:26 17 June 2013

water would crystallize in this period. Therefore, this phase is crucial to obtain frozen food
products of high quality. Fellows(96) stated that the time taken for the temperature of a food
item to pass through this period determined both the number and the size of ice crystals.
Hence, minimizing the elapsed time of the phase-change period contributes to optimizing
the quality of frozen food products. When power ultrasound was applied in this period, the
violent sonic agitation elevated the rate of removing latent heat, thus reducing the elapsed
time for phase change. Sonication is thought to improve both the rate of nucleation and the
rate of crystal growth by the disruption of preexisting nuclei thus increasing the number of
nuclei present. Fennema et al.(95) demonstrated that vigorous stirring arising from acoustic
cavitation encouraged formation of small crystals. This observation could be attributed to
both an increase in the rate of heat removal, so that the temperature decreases to a point
more favorable for nucleation, and to fragmention of growing crystals.
Delgado and Sun(94) studied the freezing rate and nucleation temperature of apple sam-
ples using ultrasound-assisted immersion freezing. Ultrasound application at 0 or −1 ◦ C for
120 s in total, with 30-s intervals, significantly improved the freezing rate up to 8%, com-
pared with immersion freezing without ultrasound. The characteristic freezing times for the
control and ultrasound treated samples were 294 and 282 s, with nucleation temperatures
of −3.1 and −2.4 ◦ C, respectively. Here, the initiation of nucleation occurred at a higher
temperature for the ultrasound treated sample, as was expected from earlier studies.(97)
Evidence of the ultrasonically induced nucleation of ice under fixed ultrasonic conditions
was reported by Chow et al.(97) for sucrose solutions containing 0–45% sucrose. Many of
the attempts to relate ultrasound with the increase in the primary nucleation of ice were
carried out in pure water systems, or in sucrose solutions, whereas the results obtained by
Delgado et al.(94) used a more complex, real food system (i.e., fruit), which also suggested
some evidence for the stimulation of primary nucleation.
Although ultrasound was found to enhance the freezing process, there has been much
interest in whether ultrasound induced changes within the structure of food materials. Sun
and Li(93) studied the microstructure of potato tissues exposed to freezing conditions under
ultrasound. Fig. 4 shows the microstructure of fresh potato cells. It can be seen that the
intact cells of fresh potato tissue were in intimate contact, although some small intercellular
voids exist. These voids are intercellular air spaces, which are common in parenchymous
tissue. The close arrangement of polyhedral potato cells endows the textural properties,
such as stiffness and crispness. Fig. 4A and B are the cryo-SEM micrographs with various
magnifications for the potato tissue that was immersion-frozen and then thawed. A compar-
ison of Fig. 4A with Fig. 4B shows that the cell shapes become less uniform after freezing
Use of Ultrasound in Food Processing 85

A B

C D
Downloaded by [University of Windsor] at 02:26 17 June 2013

Figure 4. Cryo-SEM micrographs for raw potato tissue (A); by immersion freezing and visualization
at low magnification (B); and by immersion freezing and visualization at high magnification (C).
Ultrasound-assisted immersion freezing at a frequency of 25 kHz and acoustic power of 15.85 W. A =
air space; D = disruption of cells; IV = void left by intracellular ice crystal; S = starch granule; W =
cell wall and membrane structure (from Sun and Li(93) ).

and thawing. During freezing, in the absence of ultrasound, intracellular and extracellular
crystallization occurs; the ice crystals formed occupy a larger volume than unfrozen water.
As a result, the cell walls and membranes undergo mechanical stresses arising from the vol-
umetric increase of ice crystals. The semi-rigid structure provides the possibility to resist
the stresses within a certain range. Under the mechanical stress, cells deform, thus losing
their original polyhedral shapes. The loss of ordered polyhedral cell arrangement and cell
deformation leads to a considerable reduction in the firmness of potato tissue. When the
semi-rigid structure cannot resist the strong stresses, the cell walls are disrupted, resulting
in drip loss and a reduction in food quality. Cell wall disruption is distinctly viewed in
Fig. 4B. Fig. 4C, with high magnification, shows large intercellular voids, which are much
larger than those air spaces observed in an intact tissue. It reveals the formation of large ice
crystals during freezing. The intercellular ice crystals enlarged the intercellular spaces in
slow freezing. Voids are also seen within the cells in Fig. 4C. This verifies that intracellular
crystallization occurred during immersion freezing. Overall, the freezing rate of the potato
samples by immersion freezing is slow and the structural damage is significant, leading to
the low quality of frozen potato.
When freezing is conducted in the presence of an ultrasonic power of 15.85 W
(Fig. 4D), the potato tissue cells pack together tightly with few intercellular spaces. The
defect in cellular structure seen in Fig. 4D is caused by fracture in sample preparation.
No cell wall disruption is observed. This suggests the presence of very small ice crystals
86 Chandrapala et al.

inside and outside the cells, as well as in the cell walls. As a result, the intercellular
spaces did not become enlarged, the plasma membrane remained close to the cell wall,
and the cell walls did not separate or rupture. The structural integrity is largely main-
tained, leading to a high-quality frozen product. At 15.85 W acoustic power, the freezing
rate of the potato sample was most rapid. Theoretically, small ice crystals are generated
and intracellular crystallization predominates in rapid freezing.(95) In addition, the vio-
lent agitation caused by high-intensity ultrasound when it passes through the medium is
believed to contribute to the formation of smaller ice crystals by fragmenting the crys-
tals formed. The structure deformation of ultrasonic-assisted freezing of potato tissue at
15.85 W acoustic power is minimal and can be regarded as acceptable since the cells remain
intact.
Under the influence of power ultrasound, rapid and homogeneous nucleation occurs.
In freezing, this would lead to fine ice crystals and a shortening of the time between the
onset of crystallization and the complete formation of ice, thus reducing structural damage
Downloaded by [University of Windsor] at 02:26 17 June 2013

to the cells. Acton and Morris(98) determined that when subjecting a liquid to ultrasound,
it is possible to modify both the nucleation and the crystal growth stages of solidification.
Compared with other nucleation methods, power ultrasound offers several advantages: it
allows adjustment of the initial nucleation temperature of the liquid, it is chemically non-
invasive, it does not require direct contact with the product to be frozen, and it does not
present legislative difficulties.(92) In general, crystallization in the presence of an ultrasonic
wave exhibits a number of features specific to the ultrasonic wave, which, for most mate-
rials, include (1) faster primary nucleation (formation of a nucleus in a previously crystal
free solution); (2) the initiation of secondary nucleation (nucleation induced by the pres-
ence of preexisting crystals); and (3) the production of smaller crystals with greater size
uniformity.(91) This may be primarily due to acoustic cavitation. The cavitation bubbles
can act as nuclei for crystal growth, and/or the nuclei already present can be fragmented
into smaller ones caused by the strong forces originating from the collapse of cavitation
bubbles.

Summary
The properties of power ultrasound are affected by the media through which it propagates.
In liquid media, extreme physical forces are generated by ultrasound that include acoustic
streaming, cavitation, shear, microjet, and shockwaves. The use of ultrasound in extrac-
tion processes delivers benefits such as an increase in extraction yields and rates, reduced
cost of processing by shortening the processing times, and increase in mass transfer by
greater penetration of solvent into cellular materials. Ultrasound has also proven to be use-
ful in formation of ice crystals during freezing of liquid foods due to acoustically generated
bubbles, which act as nuclei for crystal growth. It is also a useful tool in the control of crys-
tallization processes and the shapes and sizes of the crystals, due to its enhancements of the
nucleation rate and crystal growth rate. The application of high-power ultrasonic vibration
in direct contact with samples is found to be a very effective procedure for vegetable/fruit
dehydration. This procedure is a useful technique for preserving foods without affecting
their main characteristics and quality.
The motivations determining trends of development of new, emerging, or future food
technologies, such as ultrasound, are those that signify responses of food science and indus-
try to the demands of the consumers. In particular, the changes in consumer lifestyles and
expectations for fresher, more natural foods are some of the driving forces for new emerging
technologies.
Use of Ultrasound in Food Processing 87

Abbreviations
f = frequency
I = ultrasound intensity
P = ultrasound power

References
1. Mason, T.J. Sonochemistry: The Uses of Ultrasound in Chemistry; The Royal Society of
Chemistry: London, 1990.
2. Leighton, T.G. An introduction to acoustic cavitation. In Ultrasound in Medicine; Duck, F.A.,
Baker, A.C., Starritt, H.C., Eds.; Institute of Physics Publishing: London, 1998; pp 199–223.
3. Ashokkumar, M.; Mason, T. Sonochemistry; Kirk-Othmer Encylcopedia of Chemical
Technology; John Wiley & Sons: New York, 2007.
4. Louisnard, O.; Gonzalez-Garcia, J. Acoustic cavitation. In Ultrasound Technologies for Food
Downloaded by [University of Windsor] at 02:26 17 June 2013

and Bioprocessing; Feng, H., Barbosa-Canovas, G.V., Weiss, J., Eds.; Springer: New York, 2011;
pp 13–65.
5. Pankaj; Ashokkumar, M. Theoretical and Experimental Sonochemistry Involving Inorganic
Systems; Springer: New York, 2011.
6. Leighton, T.G. Bioeffects of ultrasound—The role of cavitation. In The Acoustic Bubble;
Academic Press: London, 1997; pp 551–590.
7. Toma, M.; Vinatoru, M.; Paniwnyk, L.; Mason, T.J. Investigation of the effects of ultrasound on
vegetal tissues during solvent extraction. Ultrason. Sonochem. 2001, 8, 137–142.
8. Vinatoru, M. An overview of the ultrasonically assisted extraction of bioactive principles from
herbs. Ultrason. Sonochem. 2001, 8, 303–313.
9. Li, H.; Pordesimo, L.; Weiss, J. High intensity ultrasound-assisted extraction of oil from
soybeans. Food Res. Int. 2004, 37, 731–738.
10. Vilkhu, K.; Mawson, R.; Simons, L.; Bates, D. Applications and opportunities for ultrasound
assisted extraction in the food industry—A review. Innov. Food Sci. Emerg. Technol. 2008, 9,
161–169.
11. Mason, T.; Riera, E.; Vercet, A.; Lopez-Buesa, P. Application of ultrasound. In Emerging
Technologies for Food Processing; Sun, D.W., Ed.; Elsevier Academic Press: San Diego, 2005;
pp 323–350.
12. Gómez-González, S.; Ruiz-Jiměnez, J.; Priego-Capote, F.; de Castro, M.D.L. Qualitative and
quantitative sugar profiling in olive fruits, leaves, and stems by gas chromatography–tandem
mass spectrometry (GC-MS/MS) after ultrasound-assisted leaching. J. Agric. Food Chem. 2010,
58, 12292–12299.
13. Mason T.; Zhao, M. Enhanced extraction of tea solids using ultrasound. Ultrasonics 1994, 32,
375–377.
14. Xia, T.; Shi, S.; Wan, X. Impact of ultrasound-assisted extraction on the chemical and sensory
quality of tea infusion. J. Food Eng. 2006, 74, 557–560.
15. Balachandran, S.; Kentish, S.E.; Mawson, R.; Ashokkumar, M. Ultrasonic enhancement of the
supercritical extraction from ginger. Ultrason. Sonochem. 2006, 13, 471–479.
16. Chemat, F.; Zill-e-Huma; Khan, M.K. Applications of ultrasound in food technology: Processing,
preservation and extraction. Ultrason. Sonochem. 2011, 18, 813–835.
17. Mason, T.J.; Chemat, F.; Vinatoru, M. The extraction of natural products using ultrasound or
microwaves. Curr. Org. Chem. 2011, 15, 237–247.
18. Vilkhu, K.; Manasseh, R.; Mawson, R.; Ashokkumar, M. Ultrasonic recovery and modification
of food ingredients. In Ultrasound Technologies for Food and Bioprocessing; Feng, H., Barbosa-
Canovas, G.V., Weiss, J., Eds.; Springer: New York, 2011; pp 345–368.
19. Rial-Otero, R. Ultrasonic-assisted extraction for the analysis of organic compounds by chro-
matographic techniques. In Ultrasound in Chemistry: Analytical Applications; Capelo-Martínez,
J.-L., Ed.; Wiley-VCH: Weinheim, Germany, 2009; pp 55–79.
88 Chandrapala et al.

20. Rouhani, S.; Alizadeh, N.; Salimi, S.; Haji-Ghasemi, T. Ultrasonic assisted extraction of natural
pigments from rhizomes of Curcuma Longa L. Prog. Color Colorants Coat. 2009, 2, 103–113.
21. Zou, Y.; Xie, C.; Fan, G.; Gu, Z.; Han, Y. Optimization of ultrasound-assisted extraction of
melanin from Auricularia auricula fruit bodies. Innov. Food Sci. Emerg. Technol. 2010, 11,
611–615.
22. Korn, M.; Das G.A.; Morte, E.S. da B.; dos Santos, D.C.M.B.; Castro, J.T.; Barbosa, J.T.P.;
Teixeira, A.P.; Fernandes, A.P.; Welz, B.; dos Santos, W.P.C.; dos Santos, E.B.G.N.; Korn, M.
Sample preparation for the determination of metals in food samples using spectroanalytical
methods—A review. Appl. Spec. Rev. 2008, 43, 67–92.
23. Padilha, C. do C. F.; de Moraes, P.M.; Garcia, L. de, A.; Pozzi, C.M.C.; Lima, G.P.P.; Valente,
J.P.S.; Jorge, A.M.A.; Padilha, P. de, M. Evaluation of Cu, Mn, and Se in vegetables using
ultrasonic extraction and GFAAS quantification. Food Anal. Methods 2011, 4, 319–325.
24. Rezić, I.; Horvat, A.J.M.; Babić, S.; Kaštelan-Macan, M. Determination of pesticides in honey
by ultrasonic solvent extraction and thin-layer chromatography. Ultrason. Sonochem. 2005, 12,
477–481.
25. Tadeo, J.L.; Sánchez-Brunete, C.; Albero, B.; García-Valcárcel, A.I. Application of ultrasound-
Downloaded by [University of Windsor] at 02:26 17 June 2013

assisted extraction to the determination of contaminants in food and soil samples. J. Chromatogr.
A 2010, 1217, 2415–2440.
26. Wu, J.; Lin, L.; Chau, F. Ultrasound-assisted extraction of ginseng saponins from ginseng roots
and cultured ginseng cells. Ultrason. Sonochem. 2007, 8, 347–352.
27. Cares, M.G.; Vargas, Y.; Gaete, L.; Sainz, J.; Alarcón, J. Ultrasonically assisted extraction of
bioactive principles from Quillaja saponaria Molina. Physics Procedia 2010, 3, 169–178.
28. Ebringerová, A.; Hromádková, Z. An overview on the application of ultrasound in extraction,
separation and purification of plant polysaccharides. Cent. Eur. J. Chem. 2010, 8, 243–257.
29. Kjartansson, G.T.; Zivanovic, S.; Kristbergsson, K.; Weiss, J. Sonication-assisted extraction
of chitin from North Atlantic shrimps (Pandalus borealis). J. Agric. Food Chem. 2006, 54,
5894–5902.
30. Kjartansson, G.T.; Zivanovic, S.; Kristbergsson, K.; Weiss, J. Sonication-assisted extraction of
chitin from shells of fresh water prawns (Macrobrachium rosenbergii). J. Agric. Food Chem.
2006, 54, 3317–3323.
31. Ebringerová, A.; Hromádková, Z. Effect of ultrasound on the extractability of corn bran
hemicelluloses. Ultrason. Sonochem. 2002, 9, 225–229.
32. Hromádková, Z.; Ebringerová, A. Ultrasonic extraction of plant materials-investigation of
hemicellulose release from buckwheat hulls. Ultrason. Sonochem. 2003, 10, 127–133.
33. Fu, C.; Tian, H.; Li, Q.; Cai, T.; Du, W. Ultrasound-assisted extraction of xyloglucan from apple
pomace. Ultrason. Sonochem. 2006, 13, 511–516.
34. Sun, J.X.; Sun, R.; Sun, X.F.; Su, Y. Fractional and physico-chemical characterization of hemi-
celluloses from ultrasonic irradiated sugarcane bagasse. Carbohydr. Res. 2004, 339, 291–300.
35. Zhang, M.; Zhang, L.; Cheung, P.C.K.; Ooi, V.E.C. Molecular weight and anti-tumor activity
of the water-soluble polysaccharides isolated by hot water and ultrasonic treatment from the
sclerotia and mycelia of Pleurotus tuber-regium. Carbohydr. Polym. 2004, 56, 123–128.
36. Ying, Z.; Han, X.; Li, J. Ultrasound-assisted extraction of polysaccharides from mulberry leaves.
Food Chem. 2011, 127, 1273–1279.
37. Nabarlatz, D.; Montané, D.; Kardosová, A.; Bekesová, S.; Hríbalová, V.; Ebringerová, A.
Almond shell xylo-oligosaccharides exhibiting immunostimulatory activity. Carbohydr. Res.
2007, 342, 1122–1128.
38. Košt’álová, Z.; Hromádková, Z.; Ebringerová, A. Isolation and characterization of pectic
polysaccharides from the seeded fruit of oil pumpkin (Cucurbita pepo L. var. Styriaca). Ind.
Crops Products 2010, 31, 370–377.
39. Hromádková, Z.; Košt’álová, Z.; Ebringerová, A. Comparison of conventional and ultrasound-
assisted extraction of phenolics-rich heteroxylans from wheat bran. Ultrason. Sonochem. 2008,
15, 1062–1068.
40. Sriroth, K.; Chollakup, R.; Chotineeranat, S.; Piyachomkwan, K.; Oates, C.G. Processing of
cassava waste for improved biomass utilization. Bioresource Technol. 2000, 71, 63–69.
Use of Ultrasound in Food Processing 89

41. Patist, A.; Bates, D. Industrial applications of high power ultrasonics. In Ultrasound Technologies
for Food and Bioprocessing; Feng, H., Barbosa-Canovas, G.V., Weiss, J., Eds.; Springer:
New York, 2011; pp 599–616.
42. Virot, M.; Tomao, V.; Bourvellec, C.L.; Renard, C.M.C.G.; Chemat, F. Towards the industrial
production of antioxidants from food processing by-products with ultrasound-assisted extraction.
Ultrason. Sonochem. 2010, 17, 1066–1074.
43. Paniwnyk, L.; Cai, H.; Albu, S.; Mason, T.J.; Cole, M.R. The enhancement and scale up of the
extraction of anti-oxidants from Rosmarinus officinalis using ultrasound. Ultrason. Sonochem.
2009, 16, 287–292.
44. Ashokkumar, M.; Sunartio, D.; Kentish, S.; Mawson, R.; Simons, L.; Vilkhu, K.; Versteeg, C.
Modification of food ingredients by ultrasound to improve functionality: A preliminary study on
a model system. Innov. Food Sci. Emerg. Technol. 2008, 9, 155–160.
45. Gribova, N.Y.; Filippenko, T.A.; Nikolaevskii, A.N.; Khizhan, E.L.; Bobyleva, O.V. Effects
of ultrasound on the extraction of antioxidants from bearberry (Arctostaphylos adans) leaves.
Pharm. Chem. J. 2008, 42, 593–595.
46. Pan, Z.; Qu, W.; Ma, H.; Atungulu, G.G.; McHugh, T.H. Continuous and pulsed ultrasound-
Downloaded by [University of Windsor] at 02:26 17 June 2013

assisted extractions of antioxidants from pomegranate peel. Ultrason. Sonochem. 2011, 18,
1249–1257.
47. Tiwari, B.K.; Patras, A.; Brunton, N.; Cullen, P.J.; O’Donnell, C.P. Effect of ultrasound
processing on anthocyanins and colour of red grape juice. Ultrason. Sonochem. 2010, 17,
598–604.
48. Wei, F.; Gao, G.–Z.; Wang, X.–F.; Dong, X.–Y.; Li, P.–P.; Hua, W.; Wang, X.; Wu, X.–M.;
Chen, H. Quantitative determination of oil content in small quantity of oilsed rape by ultrasound-
assisted extraction combined with gas chromatography. Ultrason. Sonochem. 2008, 15, 938–942.
49. Lou, Z.; Wang, H.; Zhang, M.; Wang, Z. Improved extraction of oil from chickpea under
ultrasound in a dynamic system. J. Food Eng. 2010, 98, 13–18.
50. Lin, J.-Y.; Zeng, Q.-X.; An, Q.; Zeng, Q.-Z.; Jian, L.-X.; Zhu, Z.-W. Ultrasonic extraction of
hempseed oil. J. Food Process Eng. 2012, 35, 76–90.
51. Chemat, S.; Lagha, A.; AitAmar, H.; Bartels, P.V.; Chemat, F. Comparison of conventional
and ultrasound-assisted extraction of carvone and limonene from caraway seeds. Flavour Frag.
J. 2004, 19, 188–195.
52. Abdullah, S.; Mudalip, S.K.A.; Shaarani, S.M.; Pi, N.A.C. Ultrasonic extraction of oil from
Monopterus albus: Effects of different ultrasonic power, solvent volume and sonication time.
J. Appl. Sci. 2010, 10, 2713–2716.
53. Zayas, J.F. Properties and quality characteristics of rennin extracted by ultrasound. Biotechnol.
Bioeng. 1986, 29, 969–975.
54. Zhang, F.-X.; Wang, B.-N. Optimization of processing parameters for the ultrasonic extraction
of goat kid rennet. Int. J. Dairy Technol. 2007, 60, 286–291.
55. Tang, D.-S.; Tian, Y.-J.; He, Y.-Z.; Li, L.; Hu, S.-Q.; Li, B. Optimization of ultrasonic-assisted
protein extraction from brewer’s spent grain. Czech. J. Food Sci. 2010, 28, 9–17.
56. Zhang, Z.; Zhang, Z.; Zhang, X.; Li, J.; Wang, Y.; Zhao, C. An ultrasound-assisted extraction
technology of almond dregs protein. Front. Agric. China 2010, 4, 69–73.
57. Mawson, R.; Gamage, M.; Terefe, M.S.; Knoerzer, K. Ultrasound in enzyme activation and inac-
tivation. In Ultrasound Technologies for Food and Bioprocessing; Feng, H., Barbosa-Canovas,
G.V., Weiss, J., Eds.; Springer: New York, 2011; pp 369–404.
58. Karki, B.; Lamsal, B.P.; Grewell, D.; Pometto, A.L., III; van Leeuwen, J.(H).; Khanal, S.K.;
Jung. S. Functional properties of soy protein isolates produced from ultrasonicated defatted soy
flakes. J. Am. Oil Chem. Soc. 2009, 86, 1021–1028.
59. Chen, Y.X.X.J.; Jian-hua, Z.Z. Effects of ultrasonification on protein extraction of defatted soy
flakes and functionality of SPI. J. South China Uni. Technol. 2003, 31, 30–36.
60. Wang, L.C. Soybean protein agglomeration: Promotion by ultrasonic treatment. J. Agric. Food
Chem. 1981, 29, 177–180.
90 Chandrapala et al.

61. Forte, T.; Gong, E.; Nichols, A.V. Interaction by sonication of C-apolipoproteins with lipid: An
electron microscopic study. Biochim. Biophys. Acta 1974, 337, 169–183.
62. Jadhav, D.; Rekha, B.N.; Parag, R.G.; Virendra, K.R. Extraction of vanillin from vanilla pods: A
comparison study of conventional Soxhlet and ultrasound assisted extraction. J. Food Eng. 2009,
93, 421–426.
63. Caldeira, I.; Pereira, R.; Cliímaco, M.C.; Belchior, A.P.; de Sousa, R.B. Improved method for
extraction of aroma compounds in aged brandies and aqueous alcoholic wood extracts using
ultrasound. Anal. Chim. Acta 2004, 513, 125–134.
64. Jofré, V.P.; Assof, M.V.; Fanzone, M.L.; Goicoechea, H.C.; Martínez, L.D.; Silva, M.F.
Optimization of ultrasound assisted-emulsification-dispersive liquid-liquid microextraction by
experimental design methodologies for the determination of sulfur compounds in wines by gas
chromatography-mass spectrometry. Anal. Chim. Acta 2010, 683, 126–135.
65. Alissandrakis, E.; Tarantilis, P.A.; Harizanis, P.C.; Polissiou, M. Evaluation of four isolation
techniques for honey aroma compounds. J. Sci. Food Agric. 2005, 85, 91–97.
66. Kimbaris, A.C.; Siatis, N.G.; Daferera, D.J.; Tarantilis, P.A.; Pappas, C.S.; Polissíou, M.G.
Comparsion of distillation and ultrasound-assisted extraction methods for the isolation of
Downloaded by [University of Windsor] at 02:26 17 June 2013

sensitive aroma compounds from garlic (Allium sativum). Ultrason. Sonochem. 2006, 13, 54–60.
67. Bund, R.K.; Pandit, A.B. Sonocrystallization: Effect on lactose recovery and crystal habit.
Ultrason. Sonochem. 2007, 14, 143–152.
68. Bund, R.K.; Pandit, A.B. Rapid lactose recovery from paneer whey using sonocrystallization: A
process optimization. Chem. Eng. Process. 2007, 46, 846–850.
69. Patel, S.R.; Murthy, V.P. Optimization of process parameters by Taguchi method in the recovery
of lactose from whey using sonocrystallization. Cryst. Res. Technol. 2010, 45, 747–752.
70. Gogate, P.R.; Mujumdar, S.; Pandit, A.B. Sonochemical reactors for waste water treatment:
Comparison using formic acid degradation as a model reaction. Adv. Environ. Res. 2003, 7,
35–39.
71. Higaki. K.; Ueno, S.; Koyano, T.; Sato, K. Effects of ultrasound irradiation on crystallization
behavior of tripalmitoylglycerol and cocoa butter. J. Am. Oil Chem. Soc. 2001, 78, 513–518.
72. Patrick, M.; Blindt, R.; Janssen J. The effect of ultrasonic intensity on the crystal structure of
plam oil. Ultrson. Sonochem. 2004, 11, 251–255.
73. Martini, S.; Suzuki, A.H.; Hartel, R.W. Effect of high intensity ultrasound on crystallization
behavior of anhydrous milk fat. J. Am. Oil Chem. Soc. 2008, 85, 621–628.
74. Suzuki, A.H.; Lee, J.; Padilla, S.G.; Martini, S. Altering functional properties of fats using power
ultrasound. J. Food Sci. 2010, 75, 208–214.
75. Sato K.; Ueno, S. Crystallization, transformation and microstructures of polymorphic fats in
colloidal dispersion states. Curr. Opin. Colloid Interface Sci. 2011, 16, 384–390.
76. Ye, Y.; Wagh, A.; Martini, S. Using high intensity ultrasound as a tool to change the functional
properties of interesterified soybean oil. J. Agric. Food Chem. 2011, 59, 10712–10722.
77. Ueno, S.; Ristic, R.I.; Higaki, K.; Sato, K. In situ studies of ultrasound stimulated fat
crystallization using synchrotron radiation. J. Phys. Chem. B 2003, 107, 4927–4935.
78. Guo, Z.; Zhang, M.; Li, H.; Wang, J.; Kougoulos, E. Effect of ultrasound on anti solvent
crystallization. J. Cryst. Growth 2005, 273, 555–558.
79. Blow, D.M.; Chayen, N.E.; Lloyd, L.F.; Saridakis, E. Control of nucleation of protein crystals.
Protein Sci. 1994, 3, 1638–1643.
80. Garcia–Ruiz, J.M. Nucleation of protein crystals. J. Struct. Biol. 2003, 142, 22–31.
81. Crespo, R.; Martin, P.M.; Gales, L.; Rocha, F.; Damas, A.M. Potential use of ultrasound to
promote protein crystallization. J. Appl. Crystallogr. 2010, 43, 1419–1425.
82. Miles, C.A.; Morley, M.J.; Rendell, M. High power ultrasonic thawing of frozen foods. J. Food
Eng. 1999, 39, 151–159.
83. Shore, D.; Woods, M.O.; Miles, C.A. Attenuation of ultrasound in post rigor bovine skeletal
muscle. Ultrasonics 1986, 24, 81–87.
84. Kissam, A.D. 1984. Acoustic thawing of frozen food. U.S. Patent No. 4464401, filed April 22,
1982.
Use of Ultrasound in Food Processing 91

85. Rosenberg, R.B.; Nesbitt, J.D.; Fejer, M.E. 1974. Method of heating frozen food using sonic or
ultrasonic wave energy. U.S. Patent No. 3846565, filed March 08, 1972.
86. Fuente-Blanco, S.; Sarabia, E.R.; Acosta-Aparicio, V.M.; Blanco-Blanco, A.; Gallego-Juárez,
J.A. Food drying process by power ultrasound. Ultrasonics 2006, 44, 523–527.
87. Fernandes, F.A.N.; Gallao, M.I.; Rodrigues, S. Effect of osmotic dehydration and ultrasound pre
treatment on cell structure: Melon dehydration. Lebensm. Wiss. Technol. 2008, 41, 604–610.
88. Azoubel, P.M.; Baima, M.A.M.; Amorim, M.R.; Oliveira, S.S.B. Effect of ultrasound on banana
cv Pacovan drying kinetics. J. Food Eng. 2010, 97, 194–198.
89. Fernandes, F.A.N.; Gallao, M.I.; Rodrigues, S. Effect of osmosis and ultrasound on pineapple
cell tissue structure during dehydration. J. Food Eng. 2009, 90, 186–190.
90. Deng, Y.; Zhao, Y. Effects of pulsed-vacuum and ultrasound on the osmodehydration kinetics
and microstructure of apples (Fuji). J. Food Eng. 2008, 85, 84–93.
91. Garcia-Perez, J.V.; Carcel, J.A.; Mulet, A. Influence of the applied acoustic energy on the drying
of carrots and lemon peel. Drying Technol. 2009, 27, 281–287.
92. Li, B.; Sun, D. Effect of power ultrasound on freezing rate during immersion freezing of potatoes.
J. Food Eng. 2002, 55, 277–282.
Downloaded by [University of Windsor] at 02:26 17 June 2013

93. Sun, D.; Li, B. Microstructural change of potato tissues frozen by ultrasound-assisted immersion
freezing. J. Food Eng. 2003, 57, 337–345.
94. Delgado, A.E.; Zheng, L.; Sun, D. Influence of ultrasound on freezing rate of immersion-frozen
apples. Food Bioprocess Technol. 2009, 2, 263–270.
95. Fennema, R.O.; Powrie, D.W.; Marth, E.H. Low Temperature Preservation of Food and Living
Matter; Marcel Dekker: New York, 1973; pp 109–368.
96. Fellows, P. Part IV: Processing by removal of heat-Freezing. Food Processing Technology—
Principles and Practice, 2nd ed.; Ellis Horward: Chichester, U.K., 2000; pp 418–440.
97. Chow, R.; Blindt, R.; Chivers, R.; Povey, M. The sonocrystallization of ice in sucrose solutions:
Primary and secondary nucleation. Ultrasonics 2003, 41, 595–604.
98. Acton E.; Morris G.J. 1992. Methods and apparatus for the control of solidification in liquids.
U.S. Patent No. WO99/20420, Filed May 05, 1992.

You might also like