Vous êtes sur la page 1sur 33

Food Bioprocess Technol (2008) 1:234

DOI 10.1007/s11947-007-0007-0

Recent Advances in the Use of High Pressure as an Effective


Processing Technique in the Food Industry
Toms Norton & Da-Wen Sun

Received: 27 May 2007 / Accepted: 17 July 2007 / Published online: 25 September 2007
# Springer Science + Business Media, LLC 2007

Abstract High pressure processing is a food processing


method which has shown great potential in the food
industry. Similar to heat treatment, high pressure processing
inactivates microorganisms, denatures proteins and extends
the shelf life of food products. But in the meantime, unlike
heat treatments, high pressure treatment can also maintain
the quality of fresh foods, with little effects on flavour and
nutritional value. Furthermore, the technique is independent
of the size, shape or composition of products. In this paper,
many aspects associated with applying high pressure as a
processing method in the food industry are reviewed,
including operating principles, effects on food quality and
safety and most recent commercial and research applications. It is hoped that this review will promote more
widespread applications of the technology to the food
industry.
Keywords High pressure . HPP . HPLT . Low temperature .
Inactivation . Enzyme . Microorganism . Shelf life .
Food quality . Food safety . Freezing . Thawing
Nomenclature
P
pressure (Pa)
T
temperature (C)

density (kg m3)

viscosity (Pa s)

T. Norton : D.-W. Sun (*)


Food Refrigeration and Computerised Food Technology Group,
University College Dublin, National University of Ireland,
Earlsfort Terrace,
Dublin 2, Ireland
e-mail: dawen.sun@ucd.ie
url: www.ucd.ie/refrig; www.ucd.ie/sun

Cp
D
k

A, B, C
CH

specific heat (W kg1 K1)


characteristic length (m)
inactivation constant
thermal conductivity (W/m1 K)
time
thermal expansion coefficient (K1)
mass of each designated food component
compression heating (C)

Subscripts
M
food medium
W
water
p
food product
pp
food product packaging
hyd_me hydraulic mechanisms in processing medium
hyd_p
hydraulic mechanisms in product
th_me
thermal conduction in processing medium
th_p
thermal conduction in food product
th_pp
thermal conduction in product packaging
in
inactivation
x, y, z
designated food component
food
composite food material

Introduction
Food processing involves synergism between different
physical processes to transform raw animal/plant materials
into consumer-ready products. Today, the food industry is
expected to prevent or reduce negative changes in food
quality over time to provide a wide variety of food rich in
colour, texture and flavour and to adapt and develop new
food processes to satisfactorily meet the requirements of a
wide demographic within different cultures. Without food

Food Bioprocess Technol (2008) 1:234

processing, these goals could not be upheld, as food could


neither be transported over long distances nor stored from
time of plenty to time of need (Lund 2002).
In the present day, consumers judge food quality based
on its sensory and nutritional characteristics (e.g. texture,
flavour, aroma, shape and colour, calorie content, vitamins
etc.), and alongside shelf life, these now determine an
individuals preference for specific products. Consequently,
retailers are reporting up to a 30% growth in fresh, chilled
and healthy food sales (Hogan et al. 2005). US sales in precut salad mixes were $1.9 billion in 2001 and increased to
$2.11 billion in 2003 (Hodge 2003). However, the recent
upsurge in demand has presented challenges to the food
industry, mainly in implementing techniques to keep food
fresher for longer, whilst offering a reasonable shelf life and
convenience and assuring food safety. Owing to recent
consumer preferences, impetus has been given to the
development of concept-driven novel technologies that
provide the required processing through non- or mildly
thermal means (Welti-Chanes et al. 2005). Accordingly,
much of the recent scientific research for the food industry
has focused on non-thermal processing techniques, with
high pressure processing (HPP) being amongst the few
experiencing great potential in commercial settings (Sun
2005).
Food safety and shelf life are often closely related to
microbial quality and other phenomena such as biochemical
reactions, enzymatic reactions and structural changes, and
thus, although often indirectly, can significantly influence
consumers perception of food quality (LeBail et al. 2003).
Physical (e.g. heating, freezing, dehydration and packaging)
and chemical (e.g. reduction of pH or use of preservatives)
preservation methods continue to be used extensively
(Manas and Pagan 2005). Conventional thermal sterilisation processes are the most commonly used methods of
food preservation and involve heat transfer from a processing medium to the slowest heating zone of a product and
subsequent cooling. Thus, although being effective mechanisms for microbial inactivation, thermal processes can
permit changes in product quality and cause off-flavour
generation, textural softening and destruction of colours
and vitamins, the extent of which is dependent on the
product being treated and the temperature gradients
between food and process boundaries. Microbial inactivation provided by HPP mainly targets cell membranes of
treated cells, but in some cases, additional damaging events
such as extensive solute loss during pressurisation, protein
denaturation and key enzyme inactivation are also required
(Manas and Pagan 2005). The multi-target ability of high
pressure (HP) has meant that in situations where its sole
employment yields unsatisfactory results, a high level of
synergism can be obtained when combined with other
processing techniques. Effective preservation has been

reported from combinations of HP with pH (Raso and


Barbosa-Canovas 2003), HP with pulsed electric fields
(Ross et al. 2003) and HP with CO2 (Spilimbergo et al.
2002). Furthermore, when used in conjunction with mildly
thermal processes, HP has been found to significantly
increase the inactivation of bacterial spores (Raso and
Barbosa-Canovas 2003).
High pressure processing is a technology that potentially
addresses many, if not all, of the most recent challenges
faced by the food industry. It can facilitate the production of
food products that have the quality of fresh foods but the
convenience and profitability associated with shelf life
extension (McClements et al. 2001). HPP has already
become a commercially implemented technology, spreading
from its origins in Japan, followed by USA and now
Europe, with worldwide take-up increasing almost exponentially since 2000 (Fig. 1a); although as of yet, this has
not been homogenous throughout the food industry. HPP
can be applied to a range of different foods, including juices
and beverages, fruits and vegetables, meat-based products
(cooked and dry ham, etc.), fish and pre-cooked dishes,
with meat and vegetables being the most popular applications (Fig. 1b). European companies presently employing
this technology include orange juice by UltiFruit; the
Pernod Ricard Company, France; and sliced ham by
Espua, Spain; fruit jams by Solofruita, Italy (UrrutiaBenet 2005). Furthermore, as evident in Table 1, a wide
variety of companies provide HPP technology to the food
industry.
High pressure processing techniques have also gained
momentum in areas of food preservation outside of
sterilisation and pasteurisation. The range of possibilities
offered by combining high pressure with low temperatures
(HPLT) has allowed the basis of a new field of HP food
applications to be formed, such as pressure-supported
freezing, thawing and subzero storage. Much work has
been conducted in the development and optimisation of
HPLT processes, and new findings regarding the phase
transitions of water, with consequential benefits for the food
industry, have recently been revealed (Urrutia-Benet et al.
2004).
High pressure research and development in different
disciplines within the food industry has been recently
reviewed by some authors (Rastogi et al. 2007; Torres and
Velazquez 2005; San Martin et al. 2002; San MartinGonzalez et al. 2006; Toepfl et al. 2006). A comprehensive
review was conducted by Rastogi et al. (2007) who, as well
as assessing many studies on the effect of HPP on enzymes
and proteins, also provided information on the successful
use of HPP, either solely or in combination with other
processing techniques. Other reviews have focused on the
effect of HPP on microorganisms and food constituents
(San Martin et al. 2002); the use of HPP in the dairy

Food Bioprocess Technol (2008) 1:234

Fig. 1 (Color online) The number of HP equipment installed in


Europe by Hyperbaric versus
a year of installment and b the
industrial sector for the installment (Urrutia-Benet 2005)

industry (OReilly et al. 2001; San Martin-Gonzalez et al.


2006; Huppertz et al. 2006); the commercial opportunities
and research challenges in HPP (Torres and Velazquez
2005); the energy efficiency of HPP (Toepfl et al. 2006)
and pressure-assisted freezing and thawing of foods
(Cheftel et al. 2002). However, no review has completed
a combined study of the modern engineering aspects of HP
technology alongside its conventional and novel uses in the
food industry. Moreover, the extensive progress made in
very recent years in non- and mildly thermal and low
temperature HPP merits a state-of-the-art review. Consequently, this study addresses many of the aspects associated
with applying high pressure as a processing method in the
food industry, from the engineering principles involved,
through food quality and safety issues, to the most recent
commercial and research applications, all of which have
seen great development in recent times.

Engineering Concepts of HPP


The governing principles of HPP are based on the
assumption that foods which experience HP in a vessel
follow the isostatic rule regardless of the size or shape of
the food. The isostatic rule states that pressure is instantaneously and uniformly transmitted throughout a sample
whether the sample is in direct contact with the pressure
medium or hermetically sealed in a flexible package.
Therefore, in contrast to thermal processing, the time
necessary for HPP should be independent of the sample
size (Rastogi et al. 2007).
The effect of HP on food chemistry and microbiology is
governed by Le Chateliers principle. This principle states
that when a system at equilibrium is disturbed, the system
then responds in a way that tends to minimise the
disturbance (Pauling 1964). In other words, HP stimulates

Food Bioprocess Technol (2008) 1:234

Table 1 Main suppliers of high pressure processing equipment and services


Company

Company specialisation

Services and/or products offered

Pressure capacity
of standard
machines (MPa)

Resato International
http://www.resato.com

This company commercialises


laboratory and industrial high
pressure hydrostatic machines
Manufactures batch presses that
pasteurize prepared ready-to-eat
foods, e.g. packaged meats
Designs and manufactures batch
presses

The company pressure shift freezing systems. They


use single shot or reciprocating intensifiers which are
suitable for one or multiple autoclave systems
Have unique pumping systems that enhance product
throughput. Continuous systems are not currently
being developed
The company has developed a system which
incorporates patent pending vessel technology. The
system that was developed exclusively for the food
processing industry from scratch
Manufacture hot, cold and warm isostatic presses

Up to 1,400

Avure Technologies Inc.,


http://www.avure.com
Elmhurst Research, Inc.,
http://www.
elmhurstresearch.com
Engineered Pressure
Systems Inc., http://www.
epsi-highpressure.com
Kobelco, http://www.
kobelco.co.jp
Mitsubishi Heavy
Industries, http://www.
mhi.co.jp
NC Hyperbaric, http://
www.nchyperbaric.com
Stansted Fluid Power
LTD. http://www.sfp-4hp.demon.co.uk
Uhde Hockdrucktechnik,
http://www.uhde-hpt.com

Manufactures laboratory and


industrial high pressure equipment
for many industries
Manufactures laboratory and
industrial high pressure equipment
for many industries
Manufactures laboratory and
industrial high pressure equipment
for many industries
European leader in manufacture of
industrial HPP equipment
Offer a full range of advanced, high
pressure equipment for research and
development applications
Uhde develop and build high
pressure processes for industry and
research purposes

some phenomena (e.g. phase transition, chemical reactivity,


change in molecular configuration, chemical reaction) that
are accompanied by a decrease in volume, but opposes
reactions that involve an increase in volume. The effects of
pressure on protein stabilisation are also governed by this
principle, i.e. the negative changes in volume with an
increase in pressure cause an equilibrium shift towards
bond formation. Alongside this, the breaking of ions is also
enhanced by HP, as this leads to a volume decrease due to
the electrostriction of water. Moreover, as hydrogen bonds
are stabilised by high pressure, as their formation involves a
volume decrease, pressure does not generally affect
covalent bonds. Consequently, HP can disrupt large
molecules of or microbial cell structures, such as enzymes,
proteins, lipids and cell membranes, and leave small
molecules such as vitamins and flavour components
unaffected (Linton and Patterson 2000).
Due to the work of compression, HPP causes temperatures to rise inside the HP vessel. This is known as
adiabatic heating and should be given due consideration

600

689

100900

Manufacture many hot and cold isostatic presses, wet


and dry-bag processes

98-686

Manufacture isostatic pressing system with large


operating temperature range as option

686

Designed a system to work with different volumes,


guaranteeing the necessary versatility to process a
wide range of products of different sizes and shapes
Single and multiple vessels with temperature control
from 20 C to +150 C. Multiple Telemetry option
and variable pressurisation times from 2s
Help in the development of plant processes from
initial testing to full scale application

600

Up to 1,400

700

during the preservation process. The value of the temperature increments in the food and pressure transmitting
medium will be different, as they depend on food
composition as well as processing temperature and pressure
and the rate of pressurisation (Otero et al. 2007a). In food
sterilisation, adiabatic heating can be used advantageously
to provide heating without the presence of sharp thermal
gradients at the process boundaries (Toepfl et al. 2006).
Knowledge of the engineering concepts of HPP has been
broadened extensively in recent times. Therefore, relevant
engineering principles that promote the capabilities of HPP
are discussed in the following.
The Mechanisms of Cellular Inactivation
The effectiveness of a food preservation technique is
primarily evaluated on the basis of its ability to eradicate
the pathogenic microorganisms that are present. Cellular
inactivation is closely associated with morphological
changes that occur within individual microbial cells during

Food Bioprocess Technol (2008) 1:234

Table 2 Mechanisms of cellular inactivation


Target microorganism

Findings

Reference

E. coli S. aureus P.
aeruginosa

Morphological changes were only noticed for the rod shaped E. coli and P. aeruginosa of which
P. aeruginosa was more pressure sensitive, whereas the S. aureus (cocci) was the most resistant
to pressure.
The pressure sensitivity of M. pneumoniae, which has no cell wall, was high compared to cell
wall gram-positive bacteria. The cell wall wasnt found to protect the bacteria and no correlation
between gram-type and pressure sensitivity was observed. However, correlation existed
between cell shape and pressure sensitivity, similar to above
Cellular morphology of L. monocytogenes was not affected when exposed to pressures of
400 MPa and membranes were perforated in small part of the population. S. typhimurium shows
morphological changes such as dimples and swellings.
Cell wall disruption occurs at 400 MPa to 500 MPA. The organelles of the cell are very sensitive
to pressure. The nuclear membrane begins to feel the affect at of 100 MPa, and at 400 MPA all
the organelles are disrupted
At 250 MPa the volume shrinkage of the cell was 15%, after compression. The volume of nonviable cells was found to be 65% after the holding time of 15 min.

Ludwig and
Schreck (1997)

M. pneumoniae

L. monocytogenes S.
typhimurium
S. cerevisiae

S. fibuligera

HPP; studies of which, as briefly reviewed by Hartmann et


al. (2006), are summarised in Table 2. From the small group
of investigations, which have thus far focussed on this area,
it is evident that cell disruption is highly specific to the
geometry of the bacteria, as opposed to its gram-type
(Ludwig and Schreck 1997; Schreck et al. 1999), although
this is disputed (Yuste et al. 2001). Moreover, the presence
of a cell wall does not mean pressure resistance is
enhanced; in fact, quite the opposite has been hypothesised
by Ludwig et al. (2002) who suggested that pressure may
induce mechanical stresses on the microbial cell wall,
which, in turn, may interact with inactivation mechanisms.
Although the above studies show strong correlations
between the physiological state of the microorganisms and
degree of pressurisation, cell disruption during processing
remains poorly understood at the fundamental level of fluid
and cell interactions (Smith et al. 2000a). Up to quite
recently, this has been quantified via a cell-wall-strength
model which presumes disruption to occur when the fluid
stresses that are imparted on a cell wall exceed some
defined threshold. This has been successfully applied to
animal cells, as these have no proper cell wall (Thomas and
Zhang 1998). Progress, however, has been slower for
microbial cells whose well-structured cell walls add
considerable complexity. As a consequence, there is a lack
of understanding and characterisation of the mechanical
properties of microbial cell walls (Smith et al. 2000a).
To appreciate the mechanical strength of microbial cells
and the factors that contribute to that strength, investigations of cell mechanical properties under periods of
pressurisation are necessary. As yeast cells are widely used
to produce intracellular bio-products of commercial interest,
experimental techniques have been employed to evaluate
their properties; for example, via micromanipulation, the

Schreck et al.
(1999)

Ritz et al. (2002)

Shimada et al.
(1993)
Perrier-Cornet et
al. (1995)

relationship between bursting force, diameter and the


relationship between force and displacement of yeast cells
have been established (Mashmoushy et al. 1998). Fortunately, yeast cell walls are structurally complex, so
experimentation may provide scope for understanding the
mechanisms of inactivation in complex microorganisms
such as Escherichia coli. In recent years, it has been found
that unless three dimensionless parameters, namely the
permeability constant, the initial thickness to radius ratio
and the initial radial stretch ratio, were found from experiments, then non-unique properties for cell walls of
biological cells could be derived (Smith et al. 1998). To
determine the cell wall properties for yeast cells using these
dimensionless parameters, Smith et al. (2000a) conducted
compression experiments. They used osmotic theory to
interpret measurements of cell volume as a function of
external osmotic pressure. Then, they quantified the effect
of osmotic pressure and cell compression rates on the
bursting force, deformation at bursting and cell diameter. To
determine the intrinsic cell wall properties and cell wall
failure criteria, the force-deformation data obtained were
used in conjunction with a finite element (FE) mechanical
model (Smith et al. 2000b). Specifically, this model
determined the mean Youngs modulus (when used in
conjunction with simple membrane theory), mean maximum von Mises stress-at-failure and mean maximum von
Mises strain-at-failure. Unfortunately, internal organelles of
the yeast cell which are also susceptible to stress were not
considered, thereby reducing the models applicability in the
area of HPP.
Hartmann and Delgado (2004) addressed this issue by
using the above information in the development of a FE
mechanical model of a yeast cell during the compression
phase of HPP (as shown in Fig. 2), which was experimen-

Food Bioprocess Technol (2008) 1:234

tion and compression rates has been shown experimentally


for other microbial species (Rademacher et al. 2002). Later,
Hartmann et al. (2006) derived a simple linear model to
explain the stress distribution on a spherical shell; although
the model assumed constant material properties, the model
still predicted the existence of heterogeneous mechanical
stresses under high hydrostatic pressure.
ThermalHydraulic Processes in HPP

Fig. 2 Finite element model of yeast cell under compression


(Hartmann and Delgado 2004)

tally validated with yeast cell volume reduction data from


Perrier-Cornet et al. (1995). Instead of using a volume loss
equation as was done in the study of Smith et al. (1998), a
reduced form of the Cauchy equation of motion represented
the mechanical behaviour of the yeast cell. Major organelles were modelled to investigate the homogeneity of the
stress distribution in the cell as well as the cell deformation
characteristics. The authors found that at 400 Mpa, the
critical effective strain upon failure of the organelles
membranes of 80% (Shimada et al. 1993) was predicted,
correlating well to experimental studies of Shimada et al.
(1993). Most notably, Hartmann and Delgado (2004)
predicted a non-homogenous (as opposed to the widely
assumed homogenous) stress distribution in the cell. In
addition, through dimensional analysis, the authors found
that the compression rates did not influence cellular
inactivation. They found that a frequency of over
700 MHz would be required for any noticeable inactivation
to occur; this frequency exceeds the feasible range of
transient pressure protocols applicable in a pulsed-HPP
system. The possibility of independence between inactiva-

As HPP often involves heat interactions and fluid flow,


thermal-hydraulic investigations, i.e. the study of thermodynamic and fluid-dynamic phenomena, have shown to be
of high importance. The thermalhydraulic processes that
occur during the HPP of both fluid and solid food systems
can be highly influential on the efficiency and effectiveness
of the overall process (Hartmann 2002; Rademacher et al.
2002). During compression/decompression phases, the
internal energy of the HP system changes, resulting in heat
transfer between the internal system and its boundaries. The
first experimental observations of fluid temperature in a HP
vessel were made by Pehl et al. (2000) who revealed a
heterogeneous temperature distribution via high-pressure
thermochromatic liquid crystals. Using the same experimental rig at room temperatures, Rademacher et al. (2002)
noted periods of forced convection during the compression/
decompression phase followed by natural convection
during the pressure holding stage. The observed temperature gradients were found to be dependent on the pressure
ramp employed. These thermalhydraulic characteristics
were also confirmed through numerical simulations by
Hartmann (2002) who noted that if food particles or
microorganisms were to be suspended in the fluid they
would undergo periodic temperature treatment with a
variation of 6 K due to a vortex motion in the pressurised
cell. Owing to the ability of the numerical simulations to
provide non-intrusive flow, thermal and concentration field
predictions, such techniques were deemed necessary in
gaining thorough understanding of the phenomena inherent
in HPP, especially when the scale-up phenomena need to be
analysed (e.g. layout and design of high pressure devices,
packages, etc.; Hartmann 2002).
An important contribution to the understanding of
thermalhydraulics in the HPP of a fluidfood system at
mild temperatures (i.e. 313 K) was made by Hartmann and
Delgado (2002). The authors used computational fluid
dynamics (CFD) and dimensional analyses to determine
the timescales of convection, conduction and bacterial
inactivation, the relative values of which contribute to the
efficiency and uniformity of conditions during HPP. These
timescales are summarised in Table 3 from which the
dependency of both convection and conduction timescales
on the geometry of the processing equipment and the

Food Bioprocess Technol (2008) 1:234

Table 3 HP-associated equations suitable for industrial application

Convection
timescale
Conduction
timescale
Inactivation
timescale
Adiabatic
temperature
rise

Fluid-food systems

Packed food systems

thyd rDh

thyd

tth

rD2 cp
l

tth

me

tin 1k

tin kk

CaTP r (1)
Y BCHZ C
CHfood CHX ACH
ABC
(2)

dT
dP

dT
dP

rDh k ; thyd
2

me

rD2 Cp k
l

; tth

CaTP r (1)

pp

Reason and
references
rD2p k
h

rpp D2pp Cppp k


; tth p
l

Y BCHZ C
CHfood CHX ACH
ABC

transport mechanisms of the fluid matrix, i.e. dynamic


viscosity and thermal conductivity, can be seen. During the
study, conductive and convective timescales were directly
compared to the inactivation timescale to provide a picture
of the thermalhydraulic states of HP vessel during
bacterial inactivation. Results of high industrial relevance
were provided as, for example; it was shown for pilot scale
systems that when processed fluids exhibit a larger
convection timescale than the inactivation timescale, intensive fluid motion and convective heat transfer resulted in
homogenous bacterial inactivation. Conversely, non-uniformities in the inactivation process were dominant when
the convection timescale was significantly smaller and the
conduction timescale was significantly larger than the
inactivation timescale. Furthermore, the simulations of
industrial-scaled systems showed greater efficiency in
bacterial inactivation as the compression heating subsisted
for greater time periods when compared to smaller
laboratory systems. As regards the HP vessel boundaries,
Otero et al. (2002a) and Hartmann et al. (2004) showed that
the thermal properties of the HP vessel boundaries have
considerable influence on the uniformity of the process, and
insulated material promoted the most effective conditions.
As well as this, the insulated vessel was found to increase
the efficiency of HPP by 40% (Hartmann et al. 2004). A
dimensionless analysis of the convective heat transfer
mechanisms in liquid foods systems under pressure was
also done by Kowalczyk and Delgado (2007a) who advised
that HP systems with a characteristic dimension of 1 m
alongside a low viscous medium should be used to avoid
heterogeneous processing of the product.
Other studies provided similar solutions to the thermal
hydraulic phenomena in HPP systems containing packaged
ultra-heat treatment (UHT) milk (Hartmann et al. 2003)
packaged enzyme mixture (Hartmann et al. 2003) solid beef
fat (Ghani and Farid 2006) and solid food analogue
material (Otero et al. 2007a), e.g. tylose with similar

(2)

rp D2p Cpp k
l

By calculating these timescales the


uniformity of HPP can be determined
(Hartmann and Delgado 2002).

Eq. 1 should be used estimate the


temperature increase for water based
foods. Eq. 2 should be used for fatty
foods (Rasanayagam et al. 2003).

properties to meat and agar with similar properties to water,


were both used. In both of the investigations of Hartmann
et al., the most significant results, revealed by validated
CFD simulations, showed strong coupling between concentrations of the surviving microorganisms and the spatial
distribution of low temperature zones within the food
package in the HP vessel. A low thermal conductive
package material was also found to improve the uniformity
of processing by preserving the elevated temperature level
within the package throughout the pressurisation phase; an
average difference of about two log reductions was found
per tenfold increase in the package thermal conductivity.
The two-dimensional CFD simulations of Otero et al.
(2007a) found that the filling ratio of the HP vessel played
a major role in process uniformity, with convective currents
having least effect on heat transfer when this ratio is large
(Fig. 3). They also showed that by anticipating the
temperature increase, which results from compression
heating, and by allowing the pressure transmitting medium
to supply the appropriate quantity of heat, the uniformity of
HPP was enhanced when both large and small sample ratios
were used (Fig. 4). More recently, Ghani and Farid (2006)
used three-dimensional CFD simulations to illustrate both
convective and conductive heat transfer in a HPP system
loaded with pieces of solid beef fat. The simulation showed
a greater adiabatic heating in the beef fat than the pressure
transmitting medium owing to the greater compression
heating coefficient used in this case.
A notable feature of the above modelling studies was
that contrasting results were possible owing to (1) the HP
systems having different operational properties or (2)
numerical modelling limitations. Therefore, different
boundary conditions have been used, and, consequently,
results between studies cannot be directly compared. For
example, in the studies of Hartmann and Delgado (2003)
and Hartmann et al. (2004), a HP vessel which permitted
the transient pressure increase as a result of the mass

Food Bioprocess Technol (2008) 1:234

Fig. 3 Temperature distribution in an HP chamber with a large filling


ratio (Otero et al. 2006)

augmentation of the inflowing pressure medium and


deformation of the packaged food, also called the indirect
HP system, was modelled. Otero et al. (2007a) and Ghani
and Farid (2006) modelled a direct system, i.e. a plungerpress which increased vessel pressure directly via the
displacement of a drive piston. Both a direct and an indirect
system are illustrated in Fig. 5. In contrast to Otero et al.
(2007a), Ghani and Farid (2006) and Hartmann et al.
(2004), Hartmann and Delgado (2003) modelled pressure

buildup to occur instantaneously in HP vessel because of


modelling limitations. The authors noted that this assumption was justified because the pressure holding time
exceeded the compression/decompression phase of HPP.
However, as investigations with a laboratory scale (0.8 l)
systems were cited in this justification, i.e. with small
convection and conduction timescales (Pehl et al. 2000),
whereas systems with much larger convection and conduction timescales were modelled (6.3 l), this must be
considered cautiously. Overall, the difference between the
boundary conditions used in these HPP modelling studies
lies in the adjustment they provided to the relative
contributions of forced and natural convection and, as a
result, their effect on temperature distribution.
It was evident from the above studies that both
temperature and velocity fields are transient during the
phase of pressure holding, as the fluid velocity distribution
influences strongly the temperature distribution and vice
versa (Otero et al. 2007a). Therefore, to accurately study
the relative contributions of forced and natural convection
to the effectiveness of HPP, it would be most beneficial to
measure velocity as well as temperature and use both to
develop a comprehensive validation in future simulation
studies.
Thermophysical Properties
Designing safe, effective and efficient HPP systems
demands the modelling of conceptual designs throughout
the range of pressures and temperatures experienced in the
food industry. One of the main difficulties when developing
or optimising these systems is the lack of knowledge about
the important thermophysical properties of food while
under pressure. However, such knowledge is important as,

Fig. 4 Temperature evolution in a big Tylose sample calculated from the model: the initial temperature of both the Tylose sample and the pressure
medium is a 40 C and b 24 C, i.e. showing the benefits of anticipating the adiabatic temperature rise (Otero et al. 2006)

10

Food Bioprocess Technol (2008) 1:234

Fig. 5 Examples of a a direct system and b an indirect systems (Urrutia-Benet 2005)

from an engineering point of view, theoretically based heat


and mass transfer models that allow the accurate prediction
of the physical history of food undergoing HP are desirable.
For example, considering the thermalhydraulic studies
reviewed above, it would not have been possible to
evaluate the relative importance of process parameters such
as the compression rate (Hartmann 2002), the size of the
HP vessel (Hartmann et al. 2003), the viscosity of the
pressure transmitting medium (Hartmann and Delgado
2002) and the process uniformity (Otero et al. 2007a) etc.
unless the physical properties of the systems fluids were
modelled as functions of pressure and temperature. For
these calculations, the thermophysical properties used
include density, viscosity, specific heat and thermal conductivity of both the pressure-transmitting medium and the
food product being processed. Of course, not all properties
have been modelled precisely, especially when limited
experimental data were available concerning the propertys
variation over the desired pressure and temperature range,
and when omitting the precise details of its dependency
would not have a large bearing on the accuracy of
simulation results, e.g. as Hartmann et al. (2003) found
when prescribing constant values for thermal conductivity
in CFD simulations (note that the variation of thermal
conductivity with pressure and temperature above freezing
point is slight as can be seen in Fig. 6.). In addition, when
HPP involves a change of phase, the ice fraction, the

enthalpy and the initial freezing point also need to be


modelled (Otero et al. 2006). Models of these properties
during HPP can be derived from (1) additive models
considering the food properties under pressure (Otero et
al. 2006); (2) in the phase change domain data at
atmospheric pressure can be shifted according to the
freezing point depression, or an experimentally observed
change, associated with the applied pressure (Denys et al.

Fig. 6 The variation in thermal conductivity of a Tylose sample with


respect to temperature and pressure (Otero et al. 2006)

Food Bioprocess Technol (2008) 1:234

1997; Hartmann et al. 2003) and (3) the physical property


of water under pressure can be multiplied by a constant
which represents the ratio of the foods physical property to
that of water at atmospheric pressure (Hartmann et al. 2003;
Ghani and Farid 2006). Another method used by Chen et al.
(2007) and Kowalczyk et al. (2005) was to firstly run twodimensional CFD simulations for a food product undergoing the HPLT process and then fit the resulting curves to
experimental data by varying the appropriate thermophysical property. A similar technique was followed by Schluter
et al. (2004) who allowed coefficients in Weibull distributions of the physical properties to vary in accordance with
the prevailing experimental conditions. The variation of
some important thermophysical properties under pressure
are discussed in the following.

Viscosity
Fluids which undergo pressurisation become more viscous
especially at subzero temperatures. Forst et al. (2000) have
published experimental data on the viscosity of water at
various temperatures as a function of pressure. Effective use
of these data permits the results obtained from viscosity
temperature equations, such as that developed by Watson et
al. (1980), to be adjusted so that the pressure experienced in
the HP system can be represented (Hartmann et al. 2004).
Many other numerical representations for viscosity of fluid
systems as a function of temperature have been published
by Seeton (2006). For liquid food systems over limited
ranges of concentration, the effect of solids concentration
on viscosity of liquid food can be described by either
exponential (Vitali and Rao 1984) or a power type of
relationship (Rao et al. 1986). The dynamic relationship
with viscosity and pressure, however, is not so well
documented. In HPP simulations of UHT milk, Hartmann
et al. (2003) considered milk to follow the same pressure
viscosity profile as that of water, represented by:

h T
hM p; T M
h p; T
1
hW T ambp W
However, owing to phenomena such as micelle disruption, the viscosity of milk during HPP cannot be explained
accurately in this way (Harte et al. 2003).

Density
The equation of state developed by Saul and Wagner
(1989), which accounts for the compressibility of pure
water under high pressure, has been used to describe
density as a function of pressure and temperature during

11

studies when convection heat transfer during HPP is being


modelled (Ghani and Farid 2006); other sources for density
data have also been used for water-like substances (Otero et
al. 2007a). As regards food, high pressure has been found
to increase the density of a food analogue by about 3.5% of
its original value for each 100-MPa increment in applied
pressure (Otero et al. 2006). Modelling compression\
decompression effects within a food sample during HPP
requires that the samples densitytemperaturepressure
relationship be taken into account. Denys et al. (2000)
measured this relationship in apple sauce and tomato paste
and regressed data to form a simple equation which they
then incorporated in their numerical heat transfer model.
When such measurements have not been possible, it was
necessary to allow the density of the food sample to vary as
a function of water density, assuming that no phase change
would occur during the HPP (Hartmann and Delgado 2003;
Ghani and Farid 2006), i.e.:

r T
rM p; T M
r p; T
2
rW T ambp W
In the phase-change domain, food density also increases
with an increase in applied pressure. Otero et al. (2006)
have shown predictions from an additive density model
under pressure to be more accurate than shifting the
atmospheric pressure density data according to the freezing
point depression. This is because shifting the data did not
take into account the increment registered in liquid water
and ice densities under pressure (Otero et al. 2006).

Specific Heat
In many foods, water substantially influences specific heat.
In addition, for matters of reducing modelling complexity,
the specific heat of the solid food components of a food
matrix can be assumed independent of temperature and
pressure (Otero et al. 2006). This means that the lower the
foods water content, the greater the difference between
predictions for the food and specific heat of water (Miles
1991). For pure water at temperatures over 0 C, increasing
the pressure causes the specific heat to decrease in an
almost linear fashion. For example, using the thermophysical data corresponding to pure water (Lemmon et al. 2005),
its specific heat at 1 C was found to decrease gradually
from 4,216 J kg1 K1 at atmospheric pressure to 3,488 J
kg1 K1 at 600 MPa. A similar gradient in the specific heat
versus pressure curve exists for all water temperatures in
the range of 0 to 120 C (Otero et al. 2002b). By assuming
that this gradient is representative of a food sample, the
specific heat of the food can then be determined as a
function of pressure. For example, in the absence of
accurate data, Ghani and Farid (2006) represented the

12

dependency of specific heat on temperature and pressure as


follows, assuming that no phase change would occur during
the HPP:

Cp M T
Cp M p; T
Cp W p; T
3
Cp W T ambp
However, it must be noted that the food should have a
high water content for this type of modelling to be accurate.
It is well known that the latent heat of fusion is reduced
under pressure and must be carefully considered when
modelling high-pressure low-temperature processes. Therefore, the apparent specific heat of foods, which includes the
contribution of the heat capacity and the latent heat of
fusion, is generally used in modelling studies. The reason
for this is that unlike the specific heat, the apparent specific
heat can be modified to account for the freezing point
depression and the reduction in latent heat of fusion via the
simple shifting approach. For more details, the reader is
referred to the articles of Otero et al. (2006) and Denys et
al. (2000).
Thermal Conductivity
In the modelling of HPP at moderate temperatures, Ghani
and Farid (2006) have followed the above methods in
describing the dependency of thermal conductivity on
temperature and pressure when no physical data for the
modelled food under pressure was available:

lM T
lM p; T
lW p; T
4
lW T ambp
In the main, thermal conductivity does not change
substantially under pressure in foods above their initial
freezing point and can even be considered constant in
modelling exercises (Hartmann et al. 2003). In the phase
change domain, both shifting the atmospheric data and
using the additive model to calculate the thermal conductivity give reasonably accurate results, as thermal conductivity shifts according to freezing point depression without
exhibiting anomalous behaviour (Otero et al. 2006).
Phase Transitions
The level of pressure imposed on a system determines the
liquidsolid phase transitions in water and food. The most
important benefits of high pressure combined with low
temperatures can be observed in the phase change diagram
of water and include (1) freezing point depression (a
minimum of 22 C at 209 MPa), (2) reduced latent heat
of fusion (from 334 kJ/kg at atmospheric pressure to
193 kJ/kg at 209 MPa), (3) a reduced change in specific
volume and (4) possibilities for the crystallisation (from

Food Bioprocess Technol (2008) 1:234

209 MPa) of higher level ice polymorphs with greater


density than water (Schluter et al. 2004). All of these
conditions are evident in Fig. 7.
Phase changes are classified according to the thermodynamic changes occurring at transition temperatures (Roos
2003). During food processing and storage, phase transitions govern the deviations in a foods physical state, with
the temperature and pressure at which they occur being
specific to the food material. As discussed by LeBail et al.
(2003), Schluter (2003) and Roos (2003), two types of
phase transitions occur in food systems, namely those of the
first and second order. In first-order transitions, the first
derivatives of the thermodynamic functions exhibit a
discontinuity in heat capacity and thermal expansion
coefficient at transition temperature (i.e. solidliquidgas
transitions). The amorphous structures of a food system,
which are formed during freezing or other forms of
processing, will undergo second-order transitions involving
no such discontinuity as, unlike first-order transitions, no
latent heat is required during the phase change; instead,
there is a step-change in the properties suffering discontinuity in the first-order transition (Roos 2003). The
existence of second-order transitions in amorphous food
structures increases the complexity of physical and chemical changes in foods (Slade and Levine 1991). The
freezing of foods gives rise to metastable, amorphous or
partially amorphous structures which exhibit time-dependent changes as they approach an equilibrium state, i.e.
crystalline (Roos 2003).
The concept of metastable states as regards the
formation of different ice types was introduced about
40 years ago (Urrutia-Benet 2005). The concept was also
recognised by Kalichevsky et al. (1995) who noted the
possibility of obtaining certain ice forms, such as ice III or
ice VI, outside their range of stability. Metastable states can
be defined as those states at which the free energy is at a
relative minimum (Schluter 2003), i.e. they correspond to a
domain in which one phase exists where another phase
would have a lower free energy. Their very existence gives
exploitable advantages to the HPLT industry. For example,
in pressure shift freezing, the presence of a metastable
supercooled liquid phase in the domains of ice I or ice III
could allow larger thermal gradients to be employed,
thereby permitting reduced processing times, and greater
amounts of ice instantaneously formed upon depressurisation (Urrutia-Benet et al. 2006). Moreover, HPLT microbial
inactivation was found to perform best in the range of
conditions corresponding to the metastable region in the
domain of ice III (Shen et al. 2005). Schluter et al. (2004)
recently provided definitions of the various metastable
phases, which, in turn, have been illustrated on the phase
change diagram of water (Fig. 8) by Urrutia-Benet et al.
(2007). Schluter et al. (2004) also showed that freezing

Food Bioprocess Technol (2008) 1:234

13

Fig. 7 The influence of pressure on the enthalpy of fusion of


ice, the specific volume changes
and the phase transition temperatures (Urrutia-Benet 2005)

within metastable states could be predicted by a onedimensional numerical heat transfer model, which used
initial freezing points obtained from an experimentally
determined phase diagram for a potato sample, illustrated in
Fig. 9. The model itself was used as a tool to give back the
corresponding values for the thermophysical properties for
each experimental condition. Doing this allowed the
authors to gain a very close fit to experimental profiles,
even when solidsolid transitions (ice Iice II) occurred. In
the comprehensive study of Schluter (2003), the authors
made some important conclusions, namely (1) as volume
changes increase concomitantly with pressure from +9% at
0.1 MPa to +13 MPa at 209 MPa, it is desirable to
pressurise the sample to the domain of ice III and as close
as possible to the triple point so that ice III has a better
chance of being formed (volume changes are 3%), (2) the

total freezing time may not be reduced when freezing to ice


III, as precooling time may be higher, (3) the degree of
supercooling is enhanced when freezing to ice III, thereby
promoting uniformity in crystal size and distribution. As
evident from Fig. 8, depending on idealised freezing or
thawing path followed in the phase transition diagram,
numerous different freezing or thawing processes can be
achieved. In fact, according to the working group of the
European project SAFE-ICE, there are in total seven
governing processes, and within this total, 13 special cases
exist (Urrutia-Benet et al. 2004). In the same study, the
authors provided clear terminology for each of these 20
processes via schematic and experimental paths and
temperature and pressure profiles; the most commercially
important of these HPLT processes will be discussed at a
later stage of this review. Supercooling was also clearly

14

Food Bioprocess Technol (2008) 1:234

Fig. 8 The metastable states


that exist on the phase diagram
of water (Urrutia-Benet 2005)

defined as the sudden temperature increase from nucleation


temperature to the initial freezing point. Table 4 summarises
the standardised nomenclature in HPP research.
Of the governing high pressure freezing and thawing
processes, those that have been modelled include subzero
cooling at high pressure, pressure-shift freezing (PSF),
pressure-assisted freezing (PAF), pressure-assisted thawing
and pressure induced thawing (PIT). Numerical modelling
can provide a clearer picture of the complex heat and mass
transfer mechanisms that govern these processes, and so it
is quickly becoming a comprehensive optimising technique
Fig. 9 The phase diagram of a
potato sample (Schluter 2003)

in freezing applications. Denys et al. (1997) were one of the


first to develop a numerical model of the conduction heat
transfer within an analogue food during PSF and PIT
processes. In their study, the thermophysical data were
shifted along the temperature melting curve according to
the prevalent pressure. Reasonable correspondence between
predictions and experimental measurements were achieved.
Later, the predictions were enhanced when the authors
correctly permitted the apparent specific heat to change as a
function of pressure (Denys et al. 2000). Many of the other
pressure-supported phase-transition modelling studies, us-

Food Bioprocess Technol (2008) 1:234

15

Table 4 Summary of HPP/HPLT terminology


Term

Definition

Reference

HPP
UHP
HHP
HP
Come-up time
Hold-time
HPLT

High pressure processing


Ultra high pressure
High hydrostatic pressure
High pressure
Time taken to pressurise the HP vessel
Time taken to maintain pressure in the HP vessel at a predefined level
High pressure low temperature

PAF

Pressure assisted freezing: an unfrozen sample is frozen after pressurization at a constant


pressure
Pressure shift freezing: a sample is frozen due to a pressure release, leading to an instantaneous
crystallization of ice, homogeneously distributed throughout the sample
Pressure induced freezing: a thawed sample can frozen by forcing to a phase transition by
pressure increase (not possible to get ice I)
Pressure assisted thawing: a sample is thawed at a constant pressure, the difference between the
sample and the bath temperature being the driving force for this process
Pressure induced thawing: a frozen sample can be forced to a phase transition from ice to liquid
water by applying pressure along the melting curve of ice I
The time span between nucleation and reaching a sample temperature (center) 5C below the
corresponding initial freezing point
The sudden temperature increase from nucleation temperature to the initial freezing point

Commonly used
Commonly used
Commonly used
Commonly used
Commonly used
Commonly used
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)
Urrutia-Benet et
al. (2004)

PSF
PIF
PAT
PIT
Plateau time or phase
transition time
Supercooling

ing conduction heat transfer models, were reviewed by


Denys et al. (2001) and Schluter et al. (2004) and will not
be covered here. Instead, the most recent contributions, all
of which include convective heat transfer from pressure
medium to the processed sample, will be reviewed.
Kowalczyk et al. (2004) were the first to model convective
heat transfer during the pressure-assisted freezing and
thawing of water. Conservation equations for phase change
were adapted to account for a compressible medium, and
alongside linearised source terms, they were solved with
CFD simulations. Contrasting heat transfer mechanisms
between freezing under atmospheric pressure and high
pressure were observed. Most notably, the authors stressed
the importance of future studies or applications to determine convection timescales for both the heating and
cooling processes and to provide correct heating parameters
during the heating phase of thawing to avoid recrystallisation. In a later study, Kowalczyk and Delgado (2007b)
found that gravity considerably influenced the shape of ice
formed under pressure, although volumetric ice formations
under low-gravity and normal conditions were not significantly different. Recently, convective and conductive heat
transfer through a tylose solution have been modelled with
the aim of determining optimum processing lengths for
semi-continuous HPLT unit, and the results indicated its
feasibility in a commercial setting (Otero et al. 2007b).
It is also worth noting the study of Ozmutlu et al. (2006)
who were the first to experimentally observe the phase
change of water under pressure. This study determined the

relative contributions of momentum and energy transfers


during the development of both ice I and ice III via particle
image velocitometry and thermography. Such encouraging
developments provide an excellent platform for the development of comprehensively validated models to gain
understanding of the physical mechanisms that govern
HPLT processes.
Developments in HPP Equipment and Processes
The general process-flow for both batch and/or semicontinuous HPP has been discussed by other authors and
will not be considered here in detail (see Hogan et al. 2005;
van den Berg et al. 2001; Mertens and Deplace 1993;
Torres and Velazquez 2005; Hjelmqwist 2005). Batch
processing is the more conventional of the two operations
and was relatively easy to implement when HPP was first
commercialised in the food industry, as hot and cold
isostatic pressing technologies could be directly adopted
from the ceramic and metal industries. For batch systems,
advances in mechanical engineering have allowed the
development of enhanced intensifier designs, advanced
opening and closing mechanisms that promote efficient
processing times and better prestressing techniques that
allow vessels to work under higher pressures with greater
fatigue resistance (van den Berg et al. 2001). A semicontinuous (or in-line) system can act as an alternative to
batch operations only when a pumpable product is being
processed. Consequently, over the years, their development

16

has been specifically aimed at the food industry. Most


notably, a semi-continuous operation promoted by many
HP system developers couples a number of pressure
systems so that most of the energy stored in a pressurised
vessel can be then used to pressurise a second vessel, thus,
saving energy and process time (van den Berg et al. 2001).
Some of the recent engineering developments and innovative concepts that have contributed to the efficiency of HPP
operations will be discussed in the following.
From a review of the patented technology, it is obvious
that scientific research has caused many of the HPP
developments in the food industry. For example, a
controlled temperature HP system has been developed
based on the adiabatic heating phenomenon (Ting and
Lonneborg 2002). The authors claimed that this system
would improve the efficacy of the pressure treatment
process by providing an insulated vessel into which the
food product could be placed. This simple concept came
about only 2 years after the research of Denys et al. (2000)
who proposed that a high level of HPP uniformity could be
achieved if the temperature increase resulting from compression was anticipated and an appropriate heat source at
the boundary of the product was then applied. As discussed
above, more recent contributions have confirmed this
hypothesis, adding more credence to the potential of this
invention (Otero et al. 2007a; Hartmann et al. 2004).
Other inventions have also been patented contemporaneously to scientific research. For example, recent studies
have observed textural changes in HP-treated vegetables to
be primarily associated with very rapid changes in
hydrostatic pressure (compression and/or decompression)
during processing, which promotes turgidity loss (TrejoAyara et al. 2007). Contemporaneously with these findings,
Ting and Anderson (2006) have developed a system and
method for decompressing a HP vessel in a controlled
manner over a selected period of time. In justifying this
invention, the authors claimed that by controlling decompression, the texture of the processed product can in turn be
controlled, and as pressure is one of the primary thermodynamic variables controlling complex biomolecular structure, controlling decompression may allow delicate
structures to remain near equilibrium. It was also suggested
that rapid decompression of a food material may cause
cellular damage due to rapid expansion of the gas that was
dissolved during pressurisation, and that slow decompression could allow gases to diffuse from structures without
cellular rupture. Although these suggestions are in line with
the scientific hypothesis of Trejo-Ayara et al. (2007), they
have yet to be proven within the scientific domain.
In batch HPP systems, the product is generally treated in
its final primary package; commonly, the food and its
package are treated together and so the entire pack remains
a secure unit until the consumer opens it. When

Food Bioprocess Technol (2008) 1:234

considering new technologies, which involve the treatment


of packaging materials, it is important to study the safety of
the material, the possible formation of compounds that
influence the odour and taste of the food and the effects of
pressure on mechanical and physical properties of the
packaging material, e.g. strength and barrier properties.
HPP requires airtight packages that can withstand a change
in volume corresponding to the compressibility of the
product (Hugas et al. 2002), as foods decrease in volume as
a function of the pressure applied, while an equal expansion
occurs on decompression. For this reason, the packaging
used for treated foods must be able to accommodate up to a
15% reduction in volume and return to its original volume
without loss of seal integrity or barrier properties. Packaging materials, which are oxygen-impermeable and opaque
to light, have been developed for keeping fresh colour and
flavour of certain HP-treated foods (Hayashi 1995). For HP
pasteurisation, a method and apparatus to store and
transport treated and untreated foods during HPP have
been developed by Hotek and Morrison (2006). In
production, the use of flexible pouches can achieve high
packing ratios; the use of semi-rigid trays is also possible,
and vacuum-packed products are ideally suited for HPP.
Miller and McLean (2006) have developed a flexible waterresistant packaging to prevent water from coming in contact
with a food product during HPP. As the size and shape of
the product will have major effects on the stacking
effectiveness of the product carrier, they must be optimised
for the most cost-effective process. This allows further
development of innovative package shapes and printing
graphics (Ting and Marshall 2002).

Recent Applications of HPP


Maintaining Food Quality Characteristics
Knowledge of the sensory and nutritional characteristics of
food products is essential for product development, quality
control, sensory evaluation and design and evaluation of
process equipment (Ahmed et al. 2003; Polydera et al.
2003). Thermal processing can often lead to quality
changes in foods such as the destruction of vitamins,
modifications to food texture and colour and the development of off-flavours. It is generally considered that HP
operations can render harmful microorganisms inactive
without having a detrimental effect on food quality (Smelt
1998). Increasing treatment pressures will generally increase microbial inactivation in shorter times, but higher
pressures may also cause greater levels of protein denaturation and other potentially detrimental changes in food
quality when compared to the unprocessed product. Yet, as
no shear forces are generated by HPP, the physical structure

Food Bioprocess Technol (2008) 1:234

17

of most high-moisture product qualities remains minimally


changed after treatment. Food characteristics which dictate
the consumers perception of food quality, and consequently the ability of HPP as a processing technology that retains
these characteristics, are reviewed in the following. Some
of the effects of HPLT process on food quality are
summarised in Table 5.
Fruit and Vegetables and Derived Products
As discussed by Cano and de Ancos (2005), the texture of
fruit and vegetable products are largely determined by the
structure of the cell wall and middle lamella. Under
pressure, the composition of these can change, as certain
cell wall enzymes are inactivated and/or structural changes
occur in the polysaccharide, lipid and protein fraction. On a
physical level, HP can disrupt the tissues morphology, cell
organelles and cell membranes (Hartmann et al. 2004).
Pressure has been shown to have a softening influence on
texture of fruits and vegetables, and tissue firmness may be
lost due to cell wall breakdown and loss of turgidity (De
Belie 2002). Trejo-Ayara et al. (2007) have found that
textural changes in raw carrots are primarily caused by loss
of turgidity induced by rapid compression and decompression. They noted that texture loss may be reduced by
turgidity manipulation of the cells or reduced by pectin
methylesterase (PME) activation during high pressure
processing given optimal conditions. In addition, they
observed loss in texture when carrots were treated with
pressures of above 300 MPa. Turgidity loss has also been
found in the cell structures of spinach, which were exposed
to a pressure level of 400 MPa for 30 min, owing to the soft
and elastic structures which characterise the cell walls; the
same was not found for tougher plant tissues such as
cauliflower (Prstamo and Arroyo 1998). Basak and

Ramaswamy (1998) found that pressure-induced textural


changes occurred in two phases, namely the textural change
due to instantaneous pressure application followed by a
gradual texture recovery or further loss during pressure
holding. In the same study, texture recovery was achieved
between 25 and 40 min for vegetable products under a
pressure of 100 MPa.
Biochemical changes also play an important role in
texture modification during HPP. PME, which is found in
plants and bacteria, de-esterifies plant cell wall pectins,
resulting in methanol and pectin with a lower degree of
methylation. In some cases, PME may enhance the texture
of fruit and vegetable products (Villarreal-Alba et al. 2004).
However, it is mostly known for inducing cloud separation
in fruit juices, making PME inactivation a prerequisite in
their processing. Moreover, the action of both PG and PME
results in the softening of plant tissues, a decrease in
viscosity, as well as cloud separation in fruit juices (Cano
and de Ancos 2005). In response to these attributes, HPP
has been used to improve or preserve the viscosity of
tomato-based products by inactivating PG whilst maintaining PME activity (Crelier et al. 2001; Fachin et al. 2002,
2004). As PME is reasonably tolerant to HP, complete
inactivation is only successful in real food samples at very
high pressures, i.e. pressures in the range 400 to 600 MPa
combined with mild heat (50 C) to accelerate PME
inactivation were advised by Nienaber and Shellhammer
(2001). Other influencing factors such as temperature, pH
and solids and protein concentrations must be considered
when pressure treating enzymes.
The colour of most fruit and vegetable products such
as jams, fruit juices and purees is generally preserved
once thresholds of temperature and/or pH are observed
(Ludikhuyze and Hendrickx 2001). For example, discolouration of broccoli juice was found after exposure to pressures

Table 5 Summary of some food quality characteristics after HPP


Product type

Treatment (MPa/C/min)

Comparison to experimental control

Reference

Orange juice
Sausages

500/35/5
500/65/5 and 15

Improved shelf-life, better consistency, lower acid loss


Better texture, improved taste, more juicy, less firm, no loss in colour

Green Beans

500 /room temp./1

Retention of colour, good firmness and extended shelf-life, showed


residual peroxidase activity
HPP showed similar reductions of vegetative cells and spores as in
heat-sterilized green beans.
Stimulated proteolysis and ultra-structural changes, tougher meat,
less juicy
Lighter colour, increased tissue firmness, shelf-life extended

Polydera et al. (2003)


Mor-Mor and Yuste
(2003)
Krebbers et al. (2002)

1000 /105/1.3
Beef

150/60/30

Salmon

200/20/10

Cheese

400/20/20

Higher yield, higher pH, reduced microbial content, less crumbly,


no colour change

Bertram et al. (2004)


Lakshmanan et al.
(2003)
Sandra et al. (2004)

18

at temperatures more than 50 C, owing to chlorophyll


degradation. However, below this temperature, pressures
of up to 800 MPa have been applied without having a
negative effect on chlorophyll (Van Loey et al. 1998). The
ability to preserve colour at high pressures is not evident in
some products, e.g. owing to polyphenol oxidase (PPO),
the colour of an onion becomes brown upon exposure to
pressure, turning browner contemporaneously with increasing pressure intensity (Butz et al. 1994). Krebbers et al.
(2003) observed an increase in colour of tomato juice when
treating the samples at 700 MPa, for 1 min at 8090 C, as
a result of compacting and homogenizing effects of the
high pressure treatment. Rodrigo et al. (2007) found that no
colour degradation of tomato appeared under combined
thermal and high pressure treatment (300700 MPa,
60 min, 65 C), and a maximum increase in colour of
8.8% was found for strawberry samples (pH 5). Thus,
recent results suggest that HPP promotes colour retention
once circumspect treatment is applied.
In many fruit and vegetable products, HPP has either no
or minor influence on flavour. Lambadarios and Zabetakis
(2002) found that HP had very little effect on strawberry
flavour compounds. Highest flavour stability was observed
when samples were treated with pressures of 200400 MPa,
and the best flavour retention was observed at 400 MPa.
Fruit juices, jams and purees all show excellent retention of
fresh like flavours for a far greater time period than that
exhibited by conventional thermal treatment under optimal
storage conditions (Ludikhuyze and Hendrickx 2001). In
fact, quite recently Baxter et al. (2005) found that HPP of
orange juice could produce a product acceptable to most
consumers even after storage for 12 weeks at temperatures
up to 10 C. On the other hand, storage at 30 C causes
900% increase in the rate of flavour deterioration (Polydera
et al. 2004).

Meat and Derived Products


As pressure bears a considerable influence on the structure
and functionality of many proteins, it consequently affects
textural, sensory and nutritive properties of meat and meatderived products (Jung et al. 2000). For meat systems, the
effectiveness of HPP depends on the characteristics
associated with the specific meat product and the intensity,
holding-time and temperature of HPP operation. Other
influencing factors include whether a meat is in a pre- or
post-rigour state, the meats pH and ionic strength, etc.
(Cheftel and Culioli 1997). Although investigations of the
effects of HPP on meat quality are limited, studies have so
far found that HP treatments can influence texture and
colour in raw, cured and battered meat systems (Jung et al.
2000; Carballo et al. 2000).

Food Bioprocess Technol (2008) 1:234

From the studies of raw meat, HPP has been shown to


tenderise meat when applied pre-rigor, but does not have a
pronounced effect on post-rigor meat at low temperatures,
with some studies even showing that HP causes meat
hardening (Jung et al. 2000). Recently, Ma and Ledward
(2004) found a massive decrease in hardness, chewiness
and cohesiveness at 200 MPa and 70 C, which they
attributed to increased enzymic activity on protein structures that have been drastically modified. At lower pressure
and temperature combinations, similar results to those
found in the literature were reported. Jung et al. (2003)
found that exposing raw meat to a high intensity of pressure
(520 MPa) for a short time (260 s) led to a decrease in the
evolution of total meat flora and a consequent delay of
growth of a week. It was then hypothesised that this delay
increases the meat maturation period, which, in turn, could
improve the meat tenderness. The authors also found meat
colour to be highly dependent on pressure intensity, as
pressures of 130 MPa improved redness, yet pressures
above 325 MPa resulted in strong discoloration, i.e. a
heightening in brown colouration. Jung et al. (2003) related
this discolouration to the increase in metmyoglobin (Fe3+)
content in the meat after pressurisation.
High pressure technology has also been employed as
a stabilising and texturising technique for meat paste
(Apichartsrangkoon and Ledward 2002; Apichartsrangkoon
2003; Jung et al. 2000). Pressure-induced changes in
protein and subsequent aggregation leads to the formation
of gels, which have better quality characteristics than those
procured through thermal means (Supavititpatana and
Apichartsrangkoon 2007). The influence of combined
pressure and heat treatment in gel formation may or may
not be synergistic, depending on the meat system under
investigation (Supavititpatana and Apichartsrangkoon
2007; Carballo et al. 2000). Nevertheless, increasing either
pressure or temperature during treatment was found to
increase gel strength, leading to a useful means of
producing meat pastes with different eating qualities
(Supavititpatana and Apichartsrangkoon 2007). When
applied to cooked sausages, Mor-Mor and Yuste (2003)
reported that HPP increased cohesiveness and reduced
firmness when compared to heat-treated sausages. They
also reported that weight loss was significantly higher in
heat-treated sausages than in HP-treated control samples.
As for changes in colour, HP-treated meat pastes became
lighter, as both the intensity of pressure and temperature
increased, thereby reducing the saleability of meat products
after processing at higher intensities (Yuste et al. 1999;
Supavititpatana and Apichartsrangkoon 2007).
For dry-cured meat products, their ability to retain
quality characteristics during HPP and throughout chilled
storage has been investigated by some authors (Rubio et al.
2007; Serra et al. 2007). Rubio et al. (2007) found that

Food Bioprocess Technol (2008) 1:234

deterioration in sensorial qualities of treated-cured ham


(500 MPa for 5 min) limited its storage time to 90 days,
although an adequate shelf life for microbial control was
found to be 210 days in the same storage conditions. Serra
et al. (2007) studied the textural and visual qualities of
pressure-treated frozen hams at different early stages in the
dry-cured process. They found that the pressurised hams
showed lower visual colour intensity than the control ones,
but did not have any significant affect on sensorial
properties of the ham. They also observed HP to increase
the ham fibrousness, which, they hypothesised, could be
useful to improve the texture of dry-cured hams with
excessive softness.
Dairy Products
As noted by Huppertz et al. (2006), although milk was the
first food to undergo HP treatment by Hite (1899), up to
now, no milk products have been commercially treated with
HP, attributed accordingly to the complex changes that milk
and derived products undergo during HP applications. The
effects of HP on milk constituents, milk properties and
bacteria that are present in milk have been comprehensively
reviewed by Huppertz et al. (2006). As well as this,
investigations into the functional improvements of milk
whey proteins promoted by HP treatment are discussed by
Lopez-Fandino (2006) and will not be covered here. Instead
of a detailed review of physiochemical and technological
changes that HP imposes on dairy products, some instances
of where the relevant functionality of dairy products, e.g.
Fig. 10 Change in the volume
of cream as a function of pressure (Huppertz et al. 2003)

19

milk and cheese, have been altered by the application of HP


technology will be discussed.
A recent finding of high importance was made by
Gervilla et al. (2001) who observed the level of free fatty
acids in ovine milk to either remain unchanged or be
reduced by HPP (500 MPa at 4, 25, 50 C), ameliorating
the effects of milk rancidity during storage. The effect on
milk fat globules was noted and seemed to be specific to
the temperature of the treatment. For example, smaller
globules were slightly increased at temperatures of both 25
and 50 C (which may have been due to the formation of
large casein aggregates; Huppertz et al. 2003), thereby
increasing milk stability, whereas at 4 C globules were
increased in size which in turn influenced the creaming
phenomenon. As seen in Fig. 10, the creaming phenomenon in raw bovine milk was recently found to be highly
dependent on the level of pressure applied, with the volume
of percentage of cream peaking at 200 MPa and reducing to
a minimum at 600 MPa (Huppertz et al. 2003). The effect
of temperature on creaming was not examined. The authors
tried to use Stokes law to explain this phenomenon, i.e. the
rate of rise of fat globules is inversely correlated with the
viscosity of the suspending medium. However, although an
increase in the milk viscosity was observed with increasing
pressure, this being attributed to the shape of the casein
micelle as well as the disruption caused to them during
treatment, the observed reduction in creaming was much
greater than that calculated from Stokes law. In addition, in
opposition to the results of Gervilla et al. (2001), no
significant HP-induced effect on milk fat globules was

20

noted. Consequently, HP-induced changes in the creaming


of milk were attributed to changes in the formation of
clusters of milk fat globules in the cold, i.e. cold
agglutination. From their findings, the authors concluded
that the use of HP presents some exciting opportunities in
the homogenisation of milk and in the development of new
milk products, as unlike traditional processing techniques,
flavour compounds are unharmed and microbial content
can be contemporaneously reduced (Huppertz et al. 2003).
In cheese making, the attributes of HP treatment are
currently being studied extensively. The most interesting,
not to mention the most economically important, investigations include those highlighting the differences between
cheese made with treated and untreated milk, the acceleration of cheese ripening and of course the reduction of
pathogenic or spoilage microbes. From studies undertaken
thus far, HP treatments at intensities greater than 200 MPa
have enhanced acid and rennet coagulation and curd
firmness times in cheese, with timescales being dependent
on the treatment temperature and pressure holding time
(San Martin-Gonzalez et al. 2007; Huppertz et al. 2005).
The main problems with using HP-treated milk for cheese,
similar to heat-treated milk (of course depending on the
type of heat treatment), are associated with the deterioration
in composition that can arise; these can even violate the
prevailing standards for cheese and are owing to the
moisture retention abilities of HP-treated milk (San
Martin-Gonzalez et al. 2007). This increased moisture
retention was suggested to be due to the formation of a finer
structural network and to the water-binding properties of
denatured -lg incorporated into the protein matrix (Needs
et al. 2000) and has also been attributed to temperature
during HP treatment (San Martin-Gonzalez et al. 2007).
Overall, HP treatment has been found to affect rennet
coagulation and other cheese-making characteristics of milk
in a fairly positive manner, although treatments could be
economically costly due to relatively long treatment time
required on expensive equipment. HP treatment can also be
conducted during the cheese-making, e.g. it was also
reported that HP treatment of Mozzarella cheese significantly accelerated the development of desirable functional
properties on melting (OReilly et al. 2002). However, the
application of HP as a pre-treatment of milk may limit the
cost of HPP (Huppertz et al. 2005).
Inactivation of Microorganisms
A primary objective of a food preservation technique is to
prevent pathogenic microorganisms from affecting the
safeness of a product. Microorganisms are resistant to
selective chemical inhibitors due to their ability to exclude
such agents from the cell, mainly by the action of the cell
membrane. However, if the cell membrane becomes

Food Bioprocess Technol (2008) 1:234

damaged, e.g. due to HP treatment, this tolerance is lost,


and the cells are vulnerable.
A secondary objective is inactivation of spoilage microorganisms to improve the shelf life of the food. Growth of
microorganisms in foods can cause spoilage by producing
unacceptable changes in taste, odour, appearance and
texture. The stage of growth of the microorganism can
have an effect on its pressure resistance, with cells in the
stationary phase being more resistant than those in the
exponential phase (McClements et al. 2001). HP treatment
is known to cause sublethal injury to microbes, which is a
particularly important consideration for any preservation
method.
Microbial inactivation by HP has been extensively
studied and has been concluded to be the result of a
combination of factors (Manas and Pagan 2005). The
primary site for pressure-induced microbial inactivation is
the cell membrane (e.g. modifications in permeability and
ion exchange; McClements et al. 2001). Microorganisms
are resistant to selective chemical inhibitors due to their
ability to exclude such agents from the cell, mainly by the
action of the cell membrane; however, if the membrane
becomes damaged, this tolerance is lost.
The ability of HP to effectively inactivate microorganisms is heavily reliant on the pressure range afforded by the
HP system, with current technology limiting commercial
HP applications to 700 MPa. Bacteria, fungi and viruses
can all be processed at pressures lower 800 MPa, the
growth and reproduction of which are severely hindered at
pressures up to 200300 MPa, with total inactivation
occurring at higher pressures. The mechanisms of microbial
inactivation including cell morphology as discussed above
and biochemical reactions and genetic mechanisms etc.,
these have been detailed by numerous authors and are not
discussed here (Hoover et al. 1989; Torres and Velazquez
2005). Instead, the following paragraphs will focus on the
influence of HP (<800 MPa) treatment on specific microorganisms.
Inactivating Bacteria
Bacteria are relatively simple, single-celled organisms and
are among the smallest free-living organisms known. The
main bacteria that cause food poisoning are Campylobacter
spp., Salmonella spp., Listeria monocytogenes, Staphyloccocus aureus, E. coli and Vibrio spp. Among these, L.
monocytogenes and S. aureus are probably the two most
intensively studied species in terms of use of HP processing. L. monocytogenes is a gram-positive rod that is an
important pathogen in acidified and other foods such as
dairy products and ready-to-eat meats. As a foodborne
bacterium, L. monocytogenes requires particular care for
processing and storage because it is moderately heat-

Food Bioprocess Technol (2008) 1:234

resistant and can grow anaerobically under refrigeration.


Nevertheless, it was shown that the use of moderate HP
treatments at mild temperatures coupled with a reasonable ripening time at refrigerating temperatures can
substantially improve cheese safety regarding L. monocytogenes (Lopez-Pedemonte et al. 2007). Alpas and
Bozoglu (2003) found that nine L. monocytogenes strains
HP-treated in fruit juice substrates developed a loss in
viability from 0.92 to 3.53 log cycles after HP treatment of
350 MPa at 25 C for 5 min. The most resistant strains (CA
and ScottA) and most sensitive strain (SLR1) differed in
viability loss by a factor of 4. However, no survivors were
detected for all of the L. monocytogenes strains when
pressurized at 350 MPa, 50 C for 5 min, suggesting this
condition as optimum for commercial inactivation. Ritz et
al. (2002) showed that although pressure treatment leads to
total inactivation of a L. monocytogenes population,
individual cells retain their morphological characteristics.
However, some physical damage was seen by the occurrence of buds on the cell surface, and membrane integrity
was lost.
S. aureus appears to have a high resistance to pressure
(Erkmen and Karatas 1997). Treating food samples using
HP can destroy both pathogenic and spoilage microorganisms; however, there is a large variation in the pressure
resistance of different bacterial strains, and the nature of the
substrate can affect the response of microorganisms to
pressure. Yet, by providing strict parameter control in
processes such as cheese making, significant reductions of
S. aureus can be obtained by means of HP coupled with a
suitable ripening time (superior to 30 days) at low storage
temperature (8 C) and short time pressure conditions
(Lopez-Pedemonte et al. 2007).
E. coli 0157:H7 also has a high barotolerance and is
considered to be an important pathogen that can cause
serious illness (Linton et al. 2001). Outbreaks of food
poisoning due to E. coli 0157:H7 have been associated with
a range of foods including ground beef (Doyle 1991), raw
and skimmed milk (Garcia-Graells et al. 1999; Linton et al.
2001); it has also been isolated from pork, lamb and poultry
(Doyle 1991; Patterson and Kilpatrick 1998). In addition,
high-acid foods such as apple cider (Besser et al. 1993),
mayonnaise (Weagent et al. 1994) and yoghurt (Morgan et
al. 1993) have also been implicated in E. coli outbreaks. It
has been found that after storage for a number of days in
refrigerated conditions, the growth of E. coli in the
pressure-treated food products is inhibited (Upmann et al.
2000). This was hypothesised to be due to pressure-induced
injury sensitising some of the bacteria to the low temperature and/or reduced oxygen conditions (Linton et al.
2000). Sublethal pressure treatment has also made E. coli
O157:H7 more susceptible to subsequent heat treatment
(Linton et al. 2000).

21

The stage of growth of the bacteria is also important in


determining pressure resistance, with cells in the stationary
phase being more resistant than those in the exponential
phase (McClements et al. 2001). Also, as discussed above,
the shape of the microorganism, as opposed to its gramtype, can significantly inactivation rate due to pressure.
Further work is required to more fully understand the
factors that can affect the response of microorganisms,
including pathogens, to pressure so that treatments can be
optimised and microbiological safety can be assured.
Inactivating Viruses
Viruses are very different from other groups of microorganisms in terms of their structure and the way in which
they function; there is also considerable diversity within the
virus family. With the exception of nucleic acid, viruses do
not have the structures that one normally associates with
living cells; they simply consist of a protein coat, called a
capsid, made up of a number of protein subunits (capsomeres) that enclose a central core of nucleic acid. Viruses
may also contain a small number of enzymes required for
the infection of host cells.
Among viruses, there is a high degree of structural
diversity, and this is reflected in a wide range of pressure
resistances (Smelt 1998). The most common human enteric
viruses are Norwalk-like viruses (SRSVs), hepatitis A,
rotavirus and human astrovirus. Complete inactivation of
suspensions of feline calicivirus (a Norwalk-like virus
surrogate), adenovirus, and adenovirus and hepatitis A can
be achieved by treatment at 275 MPa for 5 min (Kingsley et
al. 2002), 400 MPa for 15 min (Wilkinson et al. 2001) and
at 450 MPa for 5 min (Kingsley et al. 2002), respectively.
In contrast, several studies have demonstrated the remarkable baroresistance of poliovirus (Nakagami et al. 1992;
Oliveira et al. 1999; Wilkinson et al. 2001; Kingsley et al.
2002). Foot and mouth disease virus was reduced by 102.9
plaque-forming units by treatment at 220 MPa for 1 h
(Kingsley et al. 2002). The mode of inactivation of viruses
by high pressure has not been fully elucidated, although the
viral envelope, when present, appears to be one target for
HP inactivation. Treatment at pressures above 300 MPa
damages the envelopes of human immunodeficiency virus
(HIV) and cytomegalovirus, preventing the binding of virus
particles to cells (Nakagami et al. 1992). Pressure can also
cause the dissociation of virus particles; depending on the
virus and the treatment conditions, pressure-induced dissociation may be fully reversible or irreversible (Da Poian et
al. 1994). High pressure can also induce minor changes in
viral structures without disassembling the whole particle
(Gaspar et al. 2002). The formation of non-infectious
particles after HP treatment has been observed for many
viruses, including rotavirus (Pontes et al. 2001), HIV

22

(Nakagami et al. 1996), lambda phage (Bradley et al. 2000)


and picornaviruses (Oliveira et al. 1999).
Inactivating Fungi
Fungi can be divided into two groups based on their
vegetative structures: unicellular fungi (yeasts) and those
producing hyphae (moulds, mushrooms, etc.). Vegetative
bacterial cells, yeasts and moulds are, in general, more
susceptible to pressure than bacterial spores and can be
inactivated using relatively low pressures.
Yeasts are simply single-celled fungi that reproduce by
budding or fission. The group includes members of the
ascomycetes and imperfect fungi. Yeasts are an important
group of spoilage microorganisms, but are generally not
food pathogens, although toxic mould growth may be a
safety concern in foods. The Saccharomyces species are
well known for their fruit-spoiling attributes. Treatment at
pressures less than 400 MPa for a few minutes is sufficient
to inactivate most yeasts, although some strains within
species have exhibited a sluggish rate at pressures of
500 MPa (Chen and Tseng 1997). It was proposed by the
same authors that the pressure resistance of yeast is closely
related to thermal resistance. Smelt (1998) reported that at
about 100 MPa, the nuclear membrane of yeasts was
affected and that at more than 400- to 600-MPa further
alteration occurred in the mitochondria and the cytoplasm.
Moulds are mycelial fungi, and many of these organisms
are important industrially, e.g. in food spoilage, food
fermentations and biodegradation processes. Pressures
between 300 and 600 MPa can inactivate most moulds
(Smelt 1998). OReilly et al. (2000) demonstrated that HP
was effective for inactivation of Penicillium roqueforti
spores in cheese systems.

Inactivating Bacterial Spores


The elimination of bacterial endospores from food probably
represents the greatest food processing and food safety
challenges to the industry. It is well established that spores
are the most pressure-resistant life forms known; in general,
only very high pressures (>800 MPa) can kill bacterial
spores around ambient temperatures. Alternatively, other
processing methods, closer to conventional methods, can be
applied in combination with HP to effectively eliminate
bacterial spores by achieving a synergistic or hurdle effect.
In particular, HP treatment at elevated temperatures (e.g.
HP treatment at up to 90 C) has been found to be very
effective (Sangsuk and Myoung 2003). Most notably,
pressure-induced inactivation of bacterial spores is markedly enhanced at temperatures of 5070 C, and perhaps
also at or below 0 C (Okazaki et al. 2000).

Food Bioprocess Technol (2008) 1:234

The most heat-resistant pathogenic bacterium is Clostridium botulinum, and spores of C. botulinum are also
among the most pressure-resistant microorganisms known.
Among other spore-forming bacteria of concern, Bacillus
cereus has been widely studied because of its anaerobic
nature and very low rate of lethality. B. cereus is a sporeforming bacteria foodborne pathogen, which is ubiquitous
in nature, and hence occurs frequently in a wide range of
raw food materials. It is recognised as a leading cause of
bacterial food poisoning, with a variety of proteinaceaous
and starchy foods being implicated (Van Opstal et al. 2004).
An alternative to using treatments combining heat and
pressure for enhanced killing of bacterial spores is to first
cause bacterial spore germination and then use HP to kill
the much more pressure-sensitive vegetative cells. Germination is the process by which a dormant spore changes
into a vegetative cell. Interestingly, bacterial spores can be
stimulated to germinate by treatment at relatively low
pressures, e.g. 50300 MPa; germinated spores can then be
killed by relatively mild heat treatments or higher pressure
treatments (Smelt 1998). Process temperatures in the range
of 80110 C in conjunction with pressures of about
600 MPa have been used to inactivate spore-forming
bacteria such as B. cereus (Van Opstal et al. 2004). A
matrix of conditions for inactivation of spores of Bacillus
and Clostridia was presented by Meyer et al. (2000). The
mode of action of HP on bacterial spores is still largely a
matter of speculation.
Reducing the Infectivity of Prions
Prions are associated with certain neurological disorders,
including bovine spongiform encephalopathy in cattle and
CreutzfeldtJakob disease in humans. In general, prions are
even more difficult to destroy than bacterial spores, with
certain prions surviving autoclaving at 134 C (Taylor
1999). Recently, it was reported that high pressure
treatment of prion-contaminated meat at 6901,200 MPa
and 121137 C reduced the infectivity of the prions
therein (Brown et al. 2003). Such treatments may have the
advantage of ensuring safety of samples without the
excessive damage that may be associated with autoclaving
alone. Knorr et al. (2006) give a detailed review on the
latest application of HP in prion inactivation.

Recent Applications of HPLT Processing


Water, a significant component of food, has many physical
and chemical properties that are significantly affected by
pressure (Otero et al. 2002a). From the phase diagram of
water, it can be seen that the freezing temperature of water

Food Bioprocess Technol (2008) 1:234

23

decreases with increasing pressure. This phenomenon


allows the achievement of novel methods for freezing and
thawing foods. The influence of this phenomenon on lowtemperature processes in the food industry will be discussed
in the following. Some of the effects of HPLT process on
food quality are summarised in Table 6.
HPLT Processes and the Phase Diagram
High-pressure low-temperature processes are split up into
three categories, namely pressure-supported freezing, pressure-supported thawing and subzero storage. The physical
mechanisms behind these processes have been detailed by
many authors and will not be discussed here. Suffice to say
that by making reference to the phase diagram of water, one
can follow the schematic path of a desired process (Fig. 8).
However, in doing this for foods, a number of concepts
must be considered. Firstly, the presence of solutes in a
food substance will cause an additional depression of its
freezing point to a level below that of water, the magnitude
of this being a function of the concentration and molecular
weight of the solutes; the liquidIce I phase transition line
for a potato was shown to lie about 3 C below, and parallel
to, that of water (Schluter 2003). Secondly, solid and liquid
metastable states exist on the phase diagram, which means
that initial process conditions must be accurately set to
obtain a desired ice polymorph when freezing or to prevent
recrystallisation to a higher ice polymorph when thawing.
Such recrystallisation can permit drastic changes in ice
volume and can negate the benefits obtained through the
nucleation of higher density ice crystals depending on the
types of ice encountered.
During pressure-assisted freezing, the ideal freezing
point lies near the triple point of liquid/ice III/ice V, i.e.
where ice III is theoretically formed and the volume
changes from liquidsolid are low (i.e. about 3%; Schluter
et al. 2004). However, below the zone of supercooling, the
areas of nucleation of both ice I and ice III coincide,
rendering the formation of ice III unpredictable; if ice I is
formed, solidsolid transition can occur quite easily

(Urrutia-Benet et al. 2004). In addition, during pressureshift freezing, it may be important to freeze at pressures
equivalent to those that support the development of ice III
to take advantage of the higher degree of supercooling
promoted by this ice type. However, in doing so, it is
important to maintain cooling temperature at a safe level,
i.e. above that required for ice III/ice I transition, because if
this is done accurately, then ice I can be successfully
formed.
In pressure-induced thawing, the experimental path does
not follow the theoretical path as explained in Fig. 8
(Urrutia-Benet 2005). Instead, when the food is pressurised,
its temperature drops due to compression work being
converted into melting energy, along the phase transition
line. If the temperature of the food is high enough, then an
ice I/liquid phase transition will occur, often close to triple
point (liquid/ice III/ice V) of the foods phase diagram,
resulting in a negative change in volume and with a
consequential decrease in pressure. Thawing can then
continue at constant pressure. Also, the melting point may
be manipulated in that pressures of up to 300 MPa with the
correct temperature combination may be used (of course
depending on the food), taking full advantage of the large
supercooling zone within ice III (Schluter et al. 2004). On
the other hand, if the initial temperature of the sample is too
low, then an ice I/ice III phase transition will occur, with
consequential volume-changing effects. After melting has
been achieved, it is important to provide enough heat so as
to avoid recrystallisation (Denys et al. 2001).
Pressure-Induced Changes and Ice Crystal Formation
Freezing has long been established as an excellent method
for preserving food products. However, damage can be
caused by the formation of ice within a food during
freezing. The ice polymorph, Ice I, typically formed by
conventional freezing practices, is less dense than liquid
water and expands as phase change takes place. As ice
crystals are formed in conventional freezing, a freezing
front develops, inducing stress as it moves across the food

Table 6 Effect of HPLT on the quality characteristics of food


Product type

Treatment (Name/MPa/C/ min)

Comparison to experimental control

Reference

Mango and peach


Chinese cabbage
Beef

PSF /200/20/NA
PSF/100 to 700/20/45
PSF/140 to 350/22/5 to 30

Otero et al. (2000)


Fuchigami et al. (1998)
Zhao et al. (1998)

Fish and shellfish


Fish

PIT /100 to 200/15/60


PIT /200/15

Reduced thermal gradients, good microstructure


Improved texture and histological structure
Decreased thawing time, no negative effects on quality,
colour, penetration force or cooking loss
Reduced drip, reduced thawing time, no colour change
Colour changes, improved texture, reduced thawing times,
reduced drip improved microbial status

Rouille et al. (2002)


Schubring et al. (2003)

24

(LeBail et al. 2003); this can damage tissues and puncture


cell walls within foods and can also lead to the loss of
nutrients and flavour. The rate of freezing and the final
temperature of the freezing process determine the size and
location of the ice crystals formed. Fast freezing rates result
in a fine ice structure owing to intensive nucleation and
formation of high numbers of small ice crystals. This may
cause less damage to the structure of a product than
conventional atmospheric freezing (Thiebaud et al. 2002).
Owing to a high degree of supercooling, pressure-supported
freezing is known for forming instantaneous and homogenous ice crystals throughout the entire volume of a food
sample (Kalichevsky et al. 1995). However, in contrast to
this benefit, HP-induced microstructural damage may also
occur, e.g. denaturation of proteins etc., the effects of which
may lead to a degradation of quality characteristics. Recent
studies which have focused on both of these aspects in
HPLT processes will be discussed in the following.
Raw Meat Products
The influence of HP on different food matrixes during
freezing has ignited a debate among the scientific community (Fernndez-Martn 2004; LeBail 2004). This has
primarily arisen from the conflict of opinions concerning
the relative contribution of small ice crystals and pressureinduced protein denaturation to the quality of raw meats
frozen under pressure-supported means. Fernndez-Martn
(2004) suggested that this argument has been confounded
by conclusions made from studies using inadequate
experimental techniques. For example, Martino et al.
(1998), using light microscopy, studied the size and
location of ice crystals formed in raw pork meat with PSF
and concluded that from a microstructural point of view,
damage to cells was minimal because of the small size of
the ice crystals and that HP-induced distortion to fibres was
not evident. An indirect microscopic approach was again
used by Molina-Garcia et al. (2004) who found no microstructural damage associated with PAF of pork meat, this
time when frozen to ice VI. However, Fernndez-Martn
(2004) vehemently argued against the use of light microscopy, suggesting that this technique could not determine the
level of understanding which was implied in the conclusions of Martino et al. (1998) and that a combination of
transmission electron microscopy (TEM) and differential
scanning calorimeter (DSC) would be more adequate. Zhu
et al. (2004) used DSC tests to show that PSF treatment at
or above 150 MPa caused considerable denaturation of
myofibrillar proteins, which may have increased the
toughness of pork muscle during the PSF process,
especially for the 150 and 200 MPa PSF samples. Futher,
using both TEM and DSC, pressure-induced changes to
myofibrillar proteins during PSF of raw pork meat were

Food Bioprocess Technol (2008) 1:234

found by Fernandez-Martin et al. (2000), which, when


coupled with the compaction caused by the formation of
ice, irreversibly reduced the water holding capacity of the
meat samples. This consequently led Fernandez-Martin et
al. (2000) to conclude that PSF was inferior to the
conventional freezing of raw meat systems. Such a strong
conclusion was noted by both Cheftel et al. (2000) and
LeBail (2004) as not fully supported, and a sensorial
evaluation of the HP-treated raw meat after cooking was
advised to further substantiate the argument (Cheftel et al.
2000; LeBail 2004).
Fruit and Vegetable Products
Few recent studies of the pressure-supported freezing of
fruits exist. Otero et al. (2000) placed the reason for texture
loss in frozen fruit on the formation of large ice crystals
during conventional freezing. They found that PSF-treated
mangoes and peaches formed small ice crystals and did not
develop any freeze-cracking. They also observed, with light
microscopy and scanning electron microscopy, a direct
relationship between the rate of freezing and cell structural
damage, with lower freezing rates showing greater structural damage than higher freezing rates (PSF). More
recently, an analysis of the effects of PSF on the water
retention and firmness of strawberries was conducted by
Van Buggenhout et al. (2006a). The authors focused the
study on examining various means of promoting the quality
characteristics of strawberries through the combination of
freezing techniques with PME/calcium or pectin infusion;
however, the tissue microstructure after freezing was not
examined. Interestingly, and as evident in Fig. 11, raw
strawberries (no infusion) treated with PSF did not have a
positive effect on the weight loss and texture (measured as
relative firmness). In fact, weight loss was much greater
from the PSF-treated strawberries (56%) than from the
conventionally frozen strawberries (40%), with both encountering an 87% loss in their original firmness. Moreover, rapidly frozen strawberries (i.e. placed in a 18 C
cryostat bath) were found to be about 7% firmer and had
about the same weight loss than the PSF-treated ones. This
shows that the influences of PSF on the textural characteristics of soft fruit transcended those presented by the
presence of small ice crystals. However, an analysis of the
effects caused by finishing the freezing process at atmospheric pressure were not assessed; these effects may have
countered the benefits obtained from the development of
small ice crystals, as observed by the same authors when
studying carrots (Van Buggenhout et al. 2006b). The
authors also found that pressure-induced thawing did not
have any beneficial effects of texture. It must also be noted
that when infused with PME/Ca and combined with PSF,
the best texture was achieved out of all treatments.

Food Bioprocess Technol (2008) 1:234

25

Fig. 11 Hardness and drip loss


(filled diamond) of strawberries
frozen under different conditions: 1 slow freezing, 2 rapid
freezing, 3 cryogenic freezing
and 4 high-pressure shift freezing. ab Means with the same
letter indicate there is no significant difference (Tukeys test: P
<0.05) between hardness results
(bar = standard deviation; Van
Buggenhout et al. 2006a)

However, the infusion of PME/Ca had also caused the


strawberries to swell up before the PSF treatment.
Studies on the quality characteristics of blanched
broccoli samples have shown that PSF treatment did not
produce major changes in the colour and flavour. Further
investigations showed that finishing the freezing outside the
HP-vessel, in this case in liquid nitrogen, enhanced the
process efficiency whilst retaining all the quality characteristics (Fernandez et al. 2006). When Van Buggenhout et al.
(2005) measured carrot firmness after overnight storage, the
PSF-treated samples were found to be twice as firm as those
that were conventionally frozen. Interestingly, after
3 months storage, this firmness deteriorated to a level
where the carrots frozen by both techniques were found to
be almost equal. Also, PSF-treated carrots underwent a
pronounced improvement in firmness when PSF was
combined with a calcium dip and thermal or HP treatment
before the freezing step. This result was attributed to a
combination of small ice crystals, owing to the high
Fig. 12 Hardness and tissue
damage (filled diamond) of
fresh carrots at different stages
of the PSF process: i pressurization to 200 MPa, ii cooling to
15 C, iii decompression, iv
after process completion at atmospheric pressure (Van
Buggenhout et al. 2006b)

freezing rate in pressure shift freezing (PSF) and a strong


calciumpectate network, as a result of pre-treatment. This
argument is convincing considering that a synergistic effect
was not seen when the pre-treatment was combined with
conventional freezing (Van Buggenhout et al. 2005). In a
later study, Van Buggenhout et al. (2006b) found that tissue
damage in a PSF-treated carrot developed mainly during
the completion period of the freezing process, which
occurred at atmospheric pressure, as evident in Fig. 12.
The synergistic effect of pre-treating the vegetable and PSF
on textural quality was again shown, substantiating the
results of the previous study.
Studies of pressure-supported freezing and subzero
cooling have been completed by Luscher et al. (2005).
They calculated true compressive stress and strain of treated
potatoes from force-deformation curves, and their results
showed that freezing to ice III enhanced the texture of the
potato. They also showed that although PSF preserved the
skeletal cell structure of the tissue, it promoted cell

26

permeability. Volume changes during freeze-thaw cycles at


a pressure of 200 MPa were noted to have a detrimental
effect on membrane permeability (Fig. 13). During subzero
cooling, membrane permeability was also adversely influenced by phase transition within the membrane itself, which
gave rise to cell lysis. Urrutia-Benet et al. (2007) analysed
the ability of pressure-supported freezing and thawing to
inactivate PPO in potatoes and also presented a textural and
microstuctural assessment after different pressuretemperature combinations. They presented evidence that supported
the use of the metastable zone of the phase diagram to
promote the retention of colour after freezing. However,
using relative fail stress and strain values, they showed that
supercooling in the metastable zone (i.e. around 12 C
below the triple point) did not provide any additional
benefits to both the texture and drip loss of the samples
when compared to tradition PSF, i.e. 200 MPa and 20 C,
followed by PIT treatment at 200 MPa. In fact, at this
traditional pressuretemperature combination, which supports the formation of ice I, overall sample texture was
better than that of samples which underwent PSF-PIT
treatments at pressuretemperature combinations of
240 MPa/28 C, 280 MPa/20 C and 280 MPa/28 C.
Other than better colour retention and slight enzyme
inactivation, the only obvious benefit presented by supercooling into the metastable zone was seen in the reduced
swelling of cells in the potato tissue owing to the
development of smaller ice crystals.
Moreover, freezing time was increased when freezing to
the metastable zone of the phase diagram; the increase in
Fig. 13 Fractional pore area
after high pressure treatments
with phase transitions. Determined ice polymorphs and the
constant treatment pressure during the freeze-thaw cycles are
given in the legend text
(Luscher et al. 2005)

Food Bioprocess Technol (2008) 1:234

applied pressure only promotes efficiency of PIT, as shown


in their previous study (Urrutia-Benet et al. 2006).
Therefore, more work needs to be done to conclusively
prove the benefits of PSF applications which supercool
within the metastable zone of ice III. For the moment, as
suggested by Schluter et al. (2004), some considerations
must be taken into account before freezing to other ice
modifications, or cooling to ice III metastable zone, such as
volume changes, overall freezing time as well as phase
transition time. As regards process efficiency, due to the
protracted precooling time, cooling to zone of ice III does
not provide any benefits for the PSF process; in contrast,
PIT efficiency is increased.

Enhancing HPP Performance


Controlling Acidity
The pH of the food is one of the main factors affecting the
growth and survival of microorganisms; all microorganisms
have a pH range in which they can grow and an optimum
pH at which they grow best. The pH of a food, if not
optimal for a particular species, can thus not only enhance
inactivation during treatment but also inhibit outgrowth of
sublethally injured cells. Bacterial spores are generally most
resistant to the direct effects of pressure treatment at neutral
pH (Smelt 1998).
The extent of pressure-induced inactivation will generally be enhanced, and recovery of sublethally injured cells

Food Bioprocess Technol (2008) 1:234

inhibited, for most bacteria, at acidic pH values (Gao et al.


2007). For example, the pressure resistance of E. coli 0157:
H7 in orange juice is dependent on the pH of the juice and
the degree of inactivation increasing as pH decreases;
survival of E. coli 0157:H7 in orange juice during storage is
also dependent on pH (Linton and Patterson 2000).
Recently, Gao et al. (2006a) found that HP treatments
increased the death rate of S. aureus in growth medium agar
with increasing pH, but by increasing beyond a pH of 6.25,
they found that the death rate decreased. Suitable pH is also
important with regards to the quality aspect of foods.
Guiavarch et al. (2005) found that an optimum pH of 7
during HP processing was needed to enhance cloud
stability in grapefruit juice.
HP treatment of foods may shift the pH of the weakly
acidic or basic food as a function of imposed pressure. For
example, it has been shown that HP increased the pH of
minced tuna fish samples (p<0.05; Ramirez-Suarez and
Morrissey 2006). In fact, for weak acids and bases, the
equilibrium A(+)+B()<=>AB is shifted towards the
reactant A(+) and B() constituents when the reaction
volume change is positive and towards the product, AB,
when the reaction volume change is negative. The results
reported on the minced tuna fish may be explained by the
sole shift of the acid equilibrium constant (pKa) of ionisable
groups of amino acid side chains under HP.
Packaging Food for HPP
To protect food from contamination by the pressureinducing medium, as well as to enhance processing
efficiency, HPP is generally applied to a product in its final
packaging. Polymers or copolymers are generally used as
package materials and are suitable for HPP, as the
associated elasticity permits adequate pressure transmission
to the food, alongside maintaining a high sealing-ability
factor. The use of copolymer multilayer packaging films are
commonplace, and their tensile strength, heat seal strength,
oxygen permeability, vapour barrier permeability and
aroma permeability during HPP have been studied (LeBail
et al. 2006; Caner et al. 2004). In the main, HPP has been
shown to have a minimal effect on the mechanical strength
of packaging, e.g. LeBail et al. (2006) showed this for
seven different composite materials. Moreover, in the same
study, HPP only affected the water vapour permeability of
low-density polyethylene, in which case, the barrier
properties of the packaging were enhanced. Furthermore,
the small influence that HPP had on oxygen permeability
barrier of copolymer packaging contrasted the large
deterioration in the oxygen barrier that was observed in
conventional sterilisation processes (Lopez-Rubio et al.
2005). Kuebel et al. (1996) have also found that HPP had a
minimal effect on the diffusion of food components into a

27

polymer packaging material during exposure to high aroma


across the membrane of various polymer constructed
packaging materials.
Despite the generality amenability of many packaging
materials to HPP, some challenges have been presented,
generally caused by package flexibility problems. For
example, Caner et al. (2004) concluded that packaging
which composed of thin metal layers may provide a
physical barrier to diffusion but, as the metal and polymer
layers have different compressibility factors, HP treatment
promotes rupture in the less elastic component, i.e. the
metal layer.
Effect of Water Activity (aw)
Water in the liquid state is essential for the existence of all
living organisms. The amount of water available for
microbial growth is generally expressed in terms of the
water activity (aw) of the system. Lowering the water
activity of a food can significantly influence the growth of
food spoilage or food-poisoning organisms that may be
present in the raw materials or introduced during processing; this is the principle of the very old method of food
preservation by drying. Reducing aw appears to protect
microbes against inactivation by HPP; however, on the
other hand, recovery of sublethally injured cells can be
inhibited by low aw (Smelt 1998). The phenomenon of
sublethal injury can lead to an overestimation of microbial
inactivation, as counts taken immediately after HP treatment can be lower than those observed after a recovery
period (Murchie et al. 2005). Consequently, the net effect of
water activity on microbial inactivation by HP treatment
may be difficult to predict.
Factors Associated with Process Operation
Increasing treatment pressure, holding time or temperature
will generally increase the number of microorganisms
inactivated (with bacterial spores being the important
exception). While many HP treatments are performed at
ambient temperature, increasing or, to a lesser extent,
decreasing temperature has been found to increase the
inactivation rate of microorganisms during HP treatment.
Temperatures above 4550 C increase the rate of inactivation of food pathogens and spoilage microorganisms
(Palou et al. 2002). The use of high temperatures for food
processing is complicated by the fact that the large steel
cylinders in which the food is held during treatment are
very slow to change in temperature, and that the food itself
can undergo a significant increase in temperature during
processing due to adiabatic heating. Moreover, if the food
contains a significant amount of fat, such as butter or
cream, the temperature rise can be large. Foods cool down

28

to their original temperature on decompression if no heat is


lost to, or gained through, the walls of the pressure vessel
during the hold time at pressure.
Temperature increases due to adiabatic compressions can
be 3 C or more per 100 Mpa, and whilst these increases in
temperature are generally transient, in some processes, use
of sample insulation may retain this heat and add to the
thermal dimension of the processing conditions (Ting and
Marshall 2002). The choice of processing temperature will
also influence the selection of suitable pressure-transmitting
media. Recently, empirical relations were developed to
estimate the temperature increase of vegetable oil, honey
and cream cheese as a function of applied pressure and
product initial temperature (Patazca et al. 2007).
There is a minimum critical pressure below which
microbial inactivation by HP will not take place regardless
of process time. Important processing parameters to be
considered are the come-up times (period necessary to
reach treatment pressure) and pressure-release times.
Obviously, long come-up times will add appreciably to
the total process time and affect the process throughput, but
these periods will also affect inactivation kinetics of
microorganisms. Therefore, consistency and control of
these times are important in the development of HPP
techniques.

Implementing HP as an Effective Processing Technology


European Regulations for HPP
Today, in most countries, food safety is tightly controlled
by regulation. While many of the factors and microorganisms that can present hazards to the consumer are
known and have been intensively studied, new and
emerging pathogens not previously regarded as problematic
continue to be identified. As already discussed, processors
must also increasingly balance the need for assurance of
food safety against consumer demand for minimally
processed products. For these reasons, emerging technologies such as HPP are of considerable interest and potential
benefit to the food industry.
However, before the implementation of new preservation
technologies, several issues need to be addressed, such as
the mechanisms of microbial resistance and adaptation to
these new technologies, the mechanisms of microbial and
enzyme inactivation, the identification of the most resistant
and relevant microorganisms in every food habitat, the role
of bacterial stress, the robustness of the technologies, the
increased safety relative to existing technologies and, last
but not least, the legislation needed to implement them
(Hugas et al. 2002).

Food Bioprocess Technol (2008) 1:234

Two regulatory attitudes towards commercialisation of


food products manufactured using HPP have emerged, i.e.
within the EU or outside. In countries outside the EU, there
is currently no specific legislation applicable to HPP
treatment. In the USA, for example, the traditional health
regulations are applied, and products treated by HPP, such
as guacamole and oysters, have already been introduced to
the market without any specific regulation.
In EU countries, however, national regulations for new
products have been replaced, in the application of the
precautionary principle, by a community regulation for
novel foods and ingredients (Regulation 258/97/EC) and
has been in force since 1997. This novel foods legislation
establishes an evaluation and licensing system that is
compulsory for new foods and new processes. HPP food
products are novel foods as they fulfill two conditions: their
history of human consumption has so far been negligible
and, secondly, a new manufacturing process has produced
them.
In July 2001, after the last meeting of the EU
commission in charge of novel foods, several decisions
were taken to simplify the regulations. Specifically, if it is
possible to show that the new product (e.g. the HP-treated
food) is substantially equivalent to a product already on the
market, then the product can be treated at a national
regulation level and will not need to comply with the
novel food regulation (Hugas et al. 2002).
All new pressure vessels to be used in the EU have to
comply with the pressure equipment directive (PED)
regulation which came into force in 2002. This regulation is
an extension of the CE safety standard already employed
in the EU and now recognised worldwide where CE
indicates conformity with mandatory European safety
requirements. As pressure vessels of all types utilise
potentially hazardous energy, the PED regulation seeks to
identify good design, good manufacturing practices and
detailed safety assessment for safe operation and maintenance of the vessels and auxiliary parts. Also in 2001, the
UK Food Standard Agency issued a statement which
removed the novel tag from HP technology once the
processed foods are those obtained from fruit or vegetable,
have a pH below 4.2, conform to HACCP criteria and that
the germination of clostridia is prevented during the shelf
life which should not be longer than 21 days.
Commercialisation of HPP
Food quality and safety are the two main driving forces
behind the choices made by todays consumers. This means
that the food industry must adopt new technologies to
enhance the safety, nutritional quality and sensory quality
of food products. HPP is a technology that can deliver on
all of these aspects. The biggest obstacle for HP systems is

Food Bioprocess Technol (2008) 1:234

the initial capital investment required, which currently


limits its application to high-value products. As with any
new technology, the commercial feasibility of HPP depends
ultimately on its business profitability. As a consequence,
HPP is currently, in many cases, employed where existing
technologies perform unsatisfactory instead of as a replacement technology (Corkindale 2006).
The production cost of a process must be lower than the
value added to the product. The value added by HP can be
measured in terms of a higher product quality, increased
product safety and a longer product shelf life. These issues
can further translate into reduced transportation, storage,
insurance and labour costs, consumer convenience and
enhanced safety. Food companies must be able to make a
realistic cost-benefit analysis of the potential rewards in
investment in HPP. The value of HP in terms of increasing
food safety assurance, in some cases, may alone be
sufficient to justify an investment. As microbiological
standards become more widely mandated and sensitive
assessment techniques become available in food production, the financial cost of unacceptable products will be
transferred back to the producer. The value of food safety is
frequently difficult to quantify before an incident; however,
as has been observed in recent pathogen contamination
events, a producers reputation or brand name may never
recover from a single food-safety-related incident. Having a
consistently safe product is now essential in the food
industry (Ting and Marshall 2002).
The actual cost of operating a HPP plant will depend on
many factors ranging from operating pressure, cycle time
and product geometry to labour skills and energy costs. As
with all equipment, the greater the utilisation the more costeffective it is. As the technology matures and producers
gain experience, lower equipment and operations costs can
be anticipated (Ting and Marshall 2002). In many sectors of
the food industry, for example, the cooked meat and ready
meals sectors, HP treatment offers a unique opportunity to
produce fresh-tasting products, which are safe and have a
desirable shelf life. The growing markets for these sectors
and the commercially successful implementation of HP by
a number of companies in these sectors suggest that the
initial investment costs may be sustainable. It appears that
consumers are willing to pay extra for new products or
products that have higher quality and are more convenient
than the existing range (Corkindale 2006).

Conclusions
High pressure processing is an industrially tested technology that offers a natural alternative for the processing of a
wide range of different food products. It is a technology
that can achieve the food safety of heat pasteurisation

29

whilst meeting consumer demand for fresher-tasting minimally processed foods. Application of HP can inactivate
microorganisms and enzymes and modify structures whilst
having little or no effects on nutritional and sensory quality
aspects of foods. The key advantages of HP applications to
food systems are the independence of size and geometry of
the sample during processing, possibilities for low temperature treatment and the availability of a waste-free
environmentally friendly technology.

References
Ahmed, J., Ramaswamy, H. S., Alli, I., & Ngadi, M. (2003). Effect of
high pressure on rheological characteristics of liquid egg.
Lebensmittel-Wissenschaft und-Technologie, 36, 517524.
Alpas, H., & Bozoglu, F. (2003). Efficiency of high pressure treatment
for destruction of Listeria monocytogenes in fruit juices. FEMS
Immunology and Medical Microbiology, 35(3), 269273.
Apichartsrangkoon, A. (2003). Effects of high pressure on rheological
properties of soy protein gels. Food Chemistry, 80(1), 5560.
Apichartsrangkoon, A., & Ledward, D. A. (2002). Dynamic viscoelastic behaviour of high pressure treated glutensoy mixtures.
Food Chemistry, 77(3), 317323.
Basak, S., & Ramaswamy, H. (1998). Effect of high pressure
processing on the texture of selected fruits and vegetables.
Journal of Texture Studies, 29, 587601.
Baxter, I. A., Easton, K., Schneebeli, K., & Whitfield, F. B. (2005).
High pressure processing of Australian navel orange juices:
Sensory analysis and volatile flavor profiling. Innovative Food
Science and Emerging Technologies, 6(4), 372387.
Bertram, H. C., Whittaker, A. K., Shorthose, W. R., Andersen, H. J.,
& Karlsson, A. H. (2004). Water characteristics in cooked beef as
influenced by ageing and high-pressure treatmentan NMR
micro imaging study. Meat Science, 66, 301306.
Besser, R. E., Lett, S. M., Weber, J. T., Doyle, M. P., Barrett, T. J.,
Wells, J. G., et al. (1993). An outbreak of diarrhoea and
hemolytic uremic syndrome from E. coli O157:H7 in freshpressed apple cider. Journal of American Medical Association,
269, 22172220.
Bradley, D. W., Hess, R. A., Tao, F., Sciaba-Lentz, L., Remaley, A. T.,
Laugharn, J. A., et al. (2000). Pressure cycling technology: a novel
approach to virus inactivation in plasma. Transfusion, 40(2), 193200.
Brown, P., Meyer, R., Cardone, F., & Pocchiari, M. (2003). Ultrahigh-pressure inactivation of prion infectivity in processed meat:
a practical method to prevent human infection. Proceedings of
the National Academy of Sciences, 100(10), 60936097.
Butz, P., Koller, W. D., Tauscher, B., & Wolf, S. (1994). Ultra-high
pressure processing of onions: Chemical and sensory changes.
Lebensmittel-Wissenschaft und-Technologie, 27(5), 463467.
Caner, C., Hernandez, R. J., & Harte, B. R. (2004). High-pressure
processing effects on the mechanical, barrier and mass transfer
properties of food packaging flexible structures: A critical review.
Packaging Technology and Science, 17(1), 2329.
Cano, M. P., & de Ancos, B. (2005). Advances in use of high pressure
to processing and preservation of plant foods. In G. V. BarbosaCanovas, M. S. Tapia, & M. P. Cano (Eds.), Novel food
processing technologies Ch 13(pp. 361373). Boca Raton, FL,
US: CRC Press.
Carballo, J., Cofrades, S., Solas, M. T., & Jimenez-Colmenero, F.
(2000). High pressure/thermal treatment of meat batters prepared
from freeze-thawed pork. Meat Science, 544, 357364.

30
Cheftel, J. C., & Culioli, J. (1997). Effects of high pressure on meat: A
review. Meat Science, 46(3), 211236.
Cheftel, J. C., Levy, J., & Dumay, E. (2000). Pressure-assisted
freezing and thawing: principles and potential applications. Food
Reviews International, 16(4), 453483.
Cheftel, J. C., Thiebaud, M., & Dumay, E. (2002). Pressure-assisted
freezing and thawing of foods: A review of recent studies. High
Pressure Research, 22(34), 601611.
Chen, C. S., & Tseng, C. W. (1997). Effect of high hydrostatic
pressure on the temperature dependence of Saccharomyces
cerevisiae and Zygosaccharomyces rouxii. Process Biochemistry,
32(4), 337343.
Chen, C. R., Zhu, S. M., Ramaswamy, H. S., Marcotte, M., & LeBail,
A. (2007). Computer simulation of high pressure cooling of pork.
Journal of Food Engineering, 79, 401409.
Corkindale, D. (2006). Technology too risky for major players. Food
Technology and Ingredients, 31(2), 5657.
Crelier, S., Robert, M. C., Claude, J., & Juillerat, M. A. (2001).
Tomato (Lycopersicon esculentum) pectin methylesterase and
polygalacturonase behaviors regarding heat- and pressure-induced inactivation. Journal of Agricultural and Food Chemistry,
49(11), 55665575.
Da Poian, A. T., Johnston, J. E., & Silva, J. L. (1994). Differences in
stability of the three components of cowpea mosaic virus:
Implications for virus assembly and disassembly. Biochemistry,
33, 83398346.
De Belie, N. (2002). Use of physico-chemical methods for assessment
of sensory changes in carrot texture and sweetness during
cooking. Journal of Texture Studies, 33, 367388.
Denys, S., VanLoey, A. M., & Hendrickx, M. E., & Tobback, P. P.
(1997). Modeling heat transfer during high-pressure freezing and
thawing. Biotechnology Progress, 13(4), 416423.
Denys, S., Van Loey, A. M., & Hendrickx, M. E. (2000). Modelling
conductive heat transfer during high-pressure thawing processes:
Determination of latent heat as a function of pressure. Biotechnology Progress, 16(3), 447455.
Denys, S., Schluter, O., Hendrickx, M. E., & Knorr, D. (2001). Effects
of high pressure on waterice transitions in foods. In M. E.
Hendrickx, & D. Knorr (Eds.), Ultra high pressure treatment of
food, Ch. 8.(pp. 214249). London, UK: Kluwer.
Doyle, M. P. (1991). Escherichia coli O157: H7 and its significance in
foods. International Journal of Food Microbiology, 13, 207216.
Erkmen, O., & Karatas, S. (1997). Effect of high hydrostatic pressure
on Staphylococcus aureus in milk. Journal of Food Engineering,
33, 257262.
Fachin, D., Van Loey, A. M., Nguyen, B. L., Verlent, I., & Indrawati
Hendrickx, M. E. (2002). Comparative study of the inactivation
kinetics of pectinmethylesterase in tomato juice and purified
form. Biotechnology Progress, 18(4), 739744.
Fachin, D., Smout, C., Verlent, I., Nguyen, B. L., Van Loey, A. M., &
Hendrickx, M. E. (2004). Inactivation kinetics of purified tomato
polygalacturonase by thermal and high-pressure processing.
Journal of Agricultural and Food Chemistry, 52(9), 26972703.
Fernandez, P. P., Prestamo, G., Otero, L., & Sanz, P. D. (2006).
Assessment of cell damage in high-pressure-shift frozen broccoli:
Comparison with market samples. European Food Research and
Technology, 224(1), 101107.
Fernndez-Martn, F. (2004). Comments to the article: High pressure
freezing and thawing of foods: A review: By LeBail A,
Chevalier D, Mussa DM, Ghoul M. International Journal of
Refrigeration 25:50413. International Journal of Refrigeration,
27(5), 567568.
Fernandez-Martin, F., Otero, L., Solas, M. T., & Sanz, P. (2000).
Protein denaturation and structural damage during high-pressureshift freezing of porcine and bovine muscle. Journal of Food
Science, 65(6), 10021008.

Food Bioprocess Technol (2008) 1:234


Forst, P., Werner, F., & Delgado, A. (2000). The viscosity of water at
high pressuresespecially at subzero degrees centigrade. Rheologica Acta, 39(6), 566573.
Fuchigami, M., Kato, N., & Teramoto, A. (1998). High-pressurefreezing effects on textural quality of Chinese cabbage. Journal
of Food Science, 63(1), 122125.
Gao, Y.-L., Ju, X.-R., & Jiang, H.-H. (2006a). Use of response surface
methodology to investigate the effect of food constituents on
Staphylococcus aureus inactivation by high pressure and mild
heat. Process Biochemistry, 41(2), 362369.
Gao, Y.-L., Ju, X.-R., Qiu, W.-F., & Jiang, H.-H. (2007). Investigation
of the effects of food constituents on Bacillus subtilis reduction
during high pressure and moderate temperature. Food Control,
18(10), 12501257.
Garcia-Graells, C., Kristel, J., Hauben, A., & Michiels, C. W. (1999).
High-pressure inactivation and sublethal injury of pressureresistant Escherichia coli mutants in fruit juices. Applied
Environmental Microbiology, 64, 15661568.
Gaspar, L. P., Silva, A. C. B., Gomes, A. M. O., Freitas, M. S., Ano
Bom, A. P. D., Schwarcz, W. D., et al. (2002). Hydrostatic
pressure induces the fusion-active state of enveloped viruses.
Journal of Biological Chemistry, 277, 84338439.
Gervilla, R., Ferragut, V., & Guamis, B. (2001). High hydrostatic
pressure effects on color and milk-fat globule of ewes milk.
Journal of Food Science, 66(6), 880885.
Ghani, A. G. A., & Farid, M. M. (2006). Numerical simulation of
solidliquid food mixture in a high pressure processing unit using
computational fluid dynamics. Journal of Food Engineering, 80
(4), 10311042.
Guiavarch, Y., Segovia, O., Hendrickx, M., & Van Loey, A. (2005).
Purification, characterization, thermal and high-pressure inactivation of a pectin methylesterase from white grapefruit (Citrus
paradisi). Innovative Food Science & Emerging Technologies, 6
(4), 363371.
Harte, F., Luedecke, L., Swanson, B., & Barbosa-Canovas, G. V.
(2003). Low-fat set yogurt made from milk subjected to
combinations of high hydrostatic pressure and thermal processing. Journal Dairy Science, 86, 10741082.
Hartmann, C. (2002). Numerical simulation of thermodynamic and
fluid-dynamic processes during the high-pressure treatment of
fluid food systems. Innovative Food Science & Emerging
Technologies, 3(1), 1118.
Hartmann, C., & Delgado, A. (2002). Numerical simulation of
convective and diffusive transport effects on a high-pressureinduced inactivation process. Biotechnology and Bioengineering,
79(1), 94104.
Hartmann, C., & Delgado, A. (2003). The influence of transport
phenomena during high-pressure processing of packed food on
the uniformity of enzyme inactivation. Biotechnology and
Bioengineering, 82(6), 725735.
Hartmann, C., & Delgado, A. (2004). Numerical simulation of the
mechanics of a yeast cell under high hydrostatic pressure.
Journal of Food Engineering, 37(7), 977987.
Hartmann, C., Delgado, A., & Szymczyk, J. (2003). Convective and
diffusive transport effects in a high pressure induced inactivation
process of packed food. Journal of Food Engineering, 59(1), 3344.
Hartmann, C., Schuhholz, J.-P., Kitsubun, P., Chapleau, N., LeBail,
A., & Delgado, A. (2004). Experimental and numerical analysis
of the thermofluiddynamics in a high-pressure autoclave. Food
Science & Emerging Technologies, 5(4), 399411.
Hartmann, C., Mathmann, K., & Delgado, A. (2006). Mechanical
stresses in cellular structures under high hydrostatic pressure.
Food Science & Emerging Technologies, 7(12), 112.
Hayashi, R. (1995). Advances in high pressure food processing
technology in Japan. Food Processing: Recent Developments,
Gaonkar AG, 9, 185195.

Food Bioprocess Technol (2008) 1:234


Hite, B. H. (1899). The effect of high pressure in the preservation of
milk. West Virginia Agricultural Experimental Station Bulletin,
58, 1535.
Hjelmqwist, J. (2005). Commercial high pressure equipment. In G. V.
Barbosa-Canovas, M. S. Tapia, & M. P. Cano (Eds.), Novel food
processing technologies, Chapter 16 (pp. 361373). Boca Raton,
FL, USA: CRC.
Hodge, K. (2003). Salads still hot after all these years. Fresh cut
magazine (pp. 2224). Yakima: Columbia Publishing and
Design, July.
Hogan, E., Kelly, A. L., & Sun, D.-W. (2005). High pressure
processing of foods: An overview. In D.-W. Sun (Eds.),
Emerging technologies for food processing, Chapter 1 (pp. 3
32). London, UK: Elsevier.
Hoover, D. G., Metrick, C., Papineau, A. M., Farkas, D. F., & Knorr,
D. (1989). Biological effects of high hydrostatic pressure on food
micro-organisms. Food Technology, 4, 399107.
Hotek, J. P., & Morrison, J. J. (2006). Method and apparatus for
material handling for a food product using high pressure
pasteurization. U.S. Patent 2006257552; (16 Nov 2006).
Hugas, M., Garriga, M., & Monfort, J. M. (2002). New mild
technologies in meat processing: High pressure as a model
technology. Meat Science, 62, 359371.
Huppertz, T., Fox, P. F., & Kelly, A. L. (2003). High pressure-induced
changes in the creaming properties of bovine milk. Innovative
Food Science & Emerging Technologies, 4(4), 349359.
Huppertz, T., Hinz, K., Zobrist, M. R., Uniacke, T., Kelly, A. L., &
Fox, P. F. (2005). Effects of high pressure treatment on the rennet
coagulation and cheese-making properties of heated milk.
Innovative Food Science & Emerging Technologies, 6(3), 279
285.
Huppertz, T., Smiddy, M. A., Upadhyay, V. K., & Kelly, A. L. (2006).
High-pressure-induced changes in bovine milk: A review.
International Journal of Dairy Technology, 59(2) 5866.
Jung, S., De Lamballerie-Anton, M., & Ghoul, M. (2000). Textural
changes in bovine meat treated with high pressure. High Pressure
Research, 19(16), 459464.
Jung, S., Ghoul, M., & de Lamballerie-Anton, M. (2003). Influence of
high pressure on the color and microbial quality of beef meat.
Lebensmittel-Wissenschaft-und-Technologie- Food Science and
Technology, 36(6), 625631.
Kalichevsky, M. T., Knorr, D., & Lillford, P. J. (1995). Potential food
applications of high-pressure effects on ice-water transitions.
Trends in Food Science & Technology, 6, 253258.
Kingsley, D. H., Hoover, D. G., Papfragkou, E., & Richards, G. P.
(2002). Inactivation of hepatitis A virus and a calicivirus by high
hydrostatic pressure. Journal of Food Protection, 65, 16051609.
Knorr, D., Heinz, V., & Buckow, R. (2006). High pressure application
for food biopolymers. Biochemica et Biophysica Acta, 1764(3),
619631.
Kowalczyk, W., & Delgado, A. (2007a). On convection phenomena
during high pressure treatment of liquid media. High Pressure
Research, 27(1), 8592.
Kowalczyk, W., & Delgado, A. (2007b). Dimensional analysis of
thermo-fluid-dynamics of high hydrostatic pressure processes
with phase transition. International Journal of Heat and Mass
Transfer, 50(1516), 30073018.
Kowalczyk, W., Hartmann, C., & Delgado, A. (2004). Modelling and
numerical simulation of convection driven high pressure induced
phase changes. International Journal of Heat and Mass Transfer,
47(5), 10791089.
Kowalczyk, W., Hartmann, C., Luscher, C., Pohl, M., Delgado, A., &
Knorr, D. (2005). Determination of thermophysical properties of
foods under high hydrostatic pressure in combined experimental
and theoretical approach. Innovative Food Science & Emerging
Technologies, 6(3), 318326.

31
Krebbers, B., Matser, A. M., Koets, M., & Van den Berg, R. W.
(2002). Quality and storage-stability of high-pressure preserved
green beans. Journal of Food Engineering, 54, 2733.
Krebbers, B., Matser, A. M., Hoogerwerf, S. W., Moezelaar, R.,
Tomassen, M., & Van den Berg, R. W. (2003). Combined highpressure and thermal treatments for processing of tomato puree:
Evaluation of microbial inactivation and quality parameters.
Innovative Food Science & Emerging Technologies, 4(4), 377
385.
Kuebel, J., Ludwig, H., Marx, H., & Tauscher, B. (1996). Diffusion of
aroma compounds into packaging films under high-pressure.
Packaging Technology and Science, 9(3), 143152.
Lakshmanan, R., Piggott, J. R., & Paterson, A. (2003). Potential
applications of high pressure for improvement in salmon quality.
Trends in Food Science & Technology, 14, 354363.
Lambadarios, E., & Zabetakis, I. (2002). Does high hydrostatic
pressure affect fruit esters? Lebensmittel-Wissenschaft-und-Technologie- Food Science and Technology, 35(4), 362366.
LeBail, A. (2004). Reply to the Letter to the Editor: by the
corresponding author A. LeBail of the Article: High pressure
freezing and thawing of foods: A review by LeBail A, Chevalier
D, Mussa DM, Ghoul M. International Journal of Refrigeration
25, 50413. International Journal Refrigeration, 27(5), 569.
LeBail, A., Boillereaux, L., Davenel, A., Hayert, M., Lucas, T., &
Monteau, J. Y. (2003). Phase transition in foods: Effect of
pressure and methods to assess or control phase transition.
Innovative Food Science & Emerging Technologies, 4(1), 1524.
LeBail, A., Hamadami, N., & Bahuaud, S. (2006). Effect of highpressure processing on the mechanical and barrier properties of
selected packagings. Packaging Technology and Science, 19(4),
237243.
Lemmon, E. W., McLinden, M. O., & Friend, D. G. (2005).
Thermophysical properties of fluid systems. In P. J. Linstron, &
W. G. Mallard (Eds.), NIST chemistry WebBook, NIST standard
reference database number 69. June 2005. National Institute of
Standards and Technology. Gaitherburg MD, 20899. Available
from http://webbook.nist.gov.
Linton, M., & Patterson, M. F. (2000). High pressure processing of
foods for microbiological safety and quality. Acta Microbiologica
et Immunologica Hungarica, 47(23), 175182.
Linton, M., McClements, J. M. J., & Patterson, M. F. (2000). The
combined effect of high pressure and storage on the heat
sensitivity of Escherichia coli 0157:H7. Innovative Food Science
& Emerging Technologies, 1(1), 3137.
Linton, M., McClements, J. M. J., & Patterson, M.F. (2001).
Inactivation of pathogenic Escherichia coli in skimmed milk
using high hydrostatic pressure. Innovative Food Science &
Emerging Technologies, 2(2), 99104.
Lopez-Fandino, R. (2006). Functional improvement of milk whey
proteins induced by high hydrostatic pressure. Critical Reviews
in Food Science and Nutrition, 46(4), 351363.
Lopez-Pedemonte, T., Roig-Sagus, A., De Lamo, S., Gervilla, R., &
Buenaventura, G. (2007). High hydrostatic pressure treatment
applied to model cheeses made from cows milk inoculated with
Staphylococcus aureus. Food Control, 18(5), 441447.
Lopez-Rubio, A., Lagaron, J. M., Hernandez-Munoz, P., Almenar, E.,
Catala, R., Gavara, R., et al. (2005). Effect of high pressure
treatments on the properties of EVOH-based food packaging
materials. Innovative Food Science & Emerging Technologies, 6
(1), 5158.
Ludikhuyze, L., & Hendrickx, M. E. (2001). Effects of high pressure
on chemical reactions related to food quality. In M. E. Hendrickx,
& D. Knorr (Eds.), Ultra high pressure treatment of food,
Chapter 6 (pp. 17185). London, UK: Kluwer.
Ludwig, H., & Schreck, C. (1997). The inactivation of vegetative
bacteria by pressure. In K. Heremans (Eds.), High pressure

32
research in bioscience and biotechnology (pp. 221224).
Leuven: Leuven University Press.
Ludwig, H., van Almsick, G., & Schreck, C. (2002). The effect of
high hydrostatic pressure on the survival of microorganisms. In
Y. Taniguchi, H. E. Stanley, & H. Ludwig (Eds.), Biological
systems under extreme conditions (pp. 239256). Berlin: Springer.
Lund, D. B. (2002). Food engineering for the 21st century. In J. WeltiChanes, G. V. Barbosa-Canovas, & J. M. Aguilera (Eds.),
Engineering and food for the 21st century. Food Preservation
Technology Series, Chapter 44 (pp. 314). Boca Raton, FL,
USA: CRC Press LLC.
Luscher, C., Schluter, O., & Knorr, D. (2005). High pressure-low
temperature processing of foods: Impact on cell membranes,
texture, color and visual appearance of potato tissue. Innovative
Food Science & Emerging Technologies, 6(1), 5971.
Ma, H. J., & Ledward, D. A. (2004). High pressure/thermal treatment
effects on the texture of beef muscle. Meat Science, 68(3), 347355.
Manas, P., & Pagan, R. (2005). Microbial inactivation by new
technologies of food preservation. Journal of Applied Microbiology,
98(6), 13871399.
Martino, M. N., Otero, L., Sanz, P. D., & Zaritzky, N. E. (1998). Size
and location of ice crystals in pork frozen by high-pressure
assisted freezing as compared to classical methods. Meat Science,
50(3), 303313.
Mashmoushy, H., Zhang, Z., & Thomas, C. R. (1998). Micromanipulation measurement of the mechanical properties of bakers
yeast cells. Biotechnology Letters, 12, 925929.
McClements, J. M. J., Patterson, M. F., & Linton, M. (2001). The
effect of growth stage and growth temperature on high
hydrostatic pressure inactivation of some psychrotrophic bacteria
in milk. Journal of Food Protection, 64(4), 514522.
Mertens, B. A., & Deplace, G. (1993). Engineering aspects of high
pressure technology in the food industry. Food Technology, 47,
164168.
Meyer, R., Cooper, K. L., Knorr, D., & Lelieveld, H. L. M. (2000).
High pressure sterilisation of foods. Food Technology, 54(11),
6772.
Miles, C. A. (1991). The thermophysical properties of frozen foods.
In W. Bald (Eds.), Food freezing: Today and tomorrow (pp. 45
65). London: Springer.
Miller, D. S., & Mclean, C. (2006). Packaging for use with high pressure
pasteurization. U.S. Patent 20060099306; (11 May 2006).
Molina-Garcia, A. D., Otero, L., Martino, M. N., Zaritzky, N. E.,
Arabas, J., Szczepek, J., et al. (2004). Ice VI freezing of meat:
Supercooling and ultrastructural studies. Meat Science, 66(3),
709718.
Mor-Mor, M., & Yuste, J. (2003). High pressure processing applied to
cooked sausage manufacture: Physical properties and sensory
analysis. Meat Science, 65(3), 11871191.
Morgan, D., Newman, C. P., Hutchinson, D. N., Walker, A. M., Rowe,
B., & Majid, F. (1993). Verotoxin producing Escherichia coli
O157:H7 infections associated with the consumption of yogurt.
Epidemiology and Infection, 111, 181187.
Murchie, L. W., Cruz-Romero, M., Kerry, J. P., Linton, M., Patterson, M.
F., Smiddy, M., et al. (2005). High pressure processing of shellfish:
A review of microbiological and other quality aspects. Innovative
Food Science & Emerging Technologies, 6(3), 257270.
Nakagami, T., Shigehisa, T., Ohmori, T., Taji, S., Hase, A., Kimura, T., et
al. (1992). Inactivation of herpes viruses by high hydrostatic
pressure. Journal of Virological Methods, 38, 255262.
Nakagami, T., Ohno, H., Shigehisa, T., Otake, T., Mori, H., Kawahata,
T., et al. (1996). Inactivation of human immunodeficiency virus
by high hydrostatic pressure. Transfusion, 36(5), 475476.
Needs, E. C., Stenning, R. A., Gill, A. L., Ferragut, V., & Rich, G. T.
(2000). High pressure treatment of milk: Effects on casein

Food Bioprocess Technol (2008) 1:234


micelle structure and on enzymic coagulation. Journal of Dairy
Research, 67, 3142.
Nienaber, U., & Shellhammer, T. H. (2001). High-pressure processing
of orange juice: Kinetics of pectinmethylesterase inactivation.
Journal of Food Science, 66(2), 328331.
Okazaki, T., Kakugawa, K., Yoneda, T., & Suzuki, K. (2000).
Inactivation behaviour of heat-resistant bacterial spores by
thermal treatments combined with high hydrostatic pressure.
Food Science Technology, 6, 204207.
Oliveira, A. C., Ishimaru, D., Gonalves, R. B., Smith, T. J., Mason,
P., S-Carvalho, D., et al. (1999). Low temperature and pressure
stability of picornaviruses: Implications for virus uncoating.
Biophysical Journal, 76, 12701279.
OReilly, C. E., OConnor, P. M., Kelly, A. L., Beresford, T.P., &
Murphy, P. M. (2000). Use of hydrostatic pressure for inactivation of microbial contaminants in cheese. Applied and Environmental Microbiology, 66, 48904896.
OReilly, C. E., Kelly, A. L., Murphy, P. M., & Beresford, T. P. (2001).
High pressure treatment: Applications in cheese manufacture and
ripening. Trends in Food Science & Technology, 12(2), 5159.
OReilly, C. E., Murphy, P. M., Kelly, A. L., Guinee, T. P., &
Beresford, T. P. (2002). The effect of high pressure treatment on
the functional and rheological properties of Mozzarella cheese.
Innovative Food Science & Emerging Technologies, 3, 39.
Otero, L., Martino, M., Zaritzky, N., Solas, M., & Sanz, P. D. (2000).
Preservation of microstructure in peach and mango during highpressure-shift freezing. Journal of Food Science, 65(3), 466470.
Otero, L., Molina, A., Ramos, A., & Sanz, P. D. (2002a). A model for
a real thermal control in high-pressure treatment of foods.
Biotechnology Progress, 18(4), 904908.
Otero, L., Molina-Garca, A. D., & Sanz, P. D. (2002b). Some
interrelated thermophysical properties of liquid water and ice I: A
user-friendly modelling review for high-pressure processing.
Critical Reviews in Food Science and Nutrition, 44(2), 339352.
Otero, L., Ousegui, A., Guignon, B., LeBail, A., & Sanz, P. D. (2006).
Evaluation of the thermophysical properties of tylose gel under
pressure in the phase change domain. Food Hydrocoll, 20(4),
449460.
Otero, L., Ramos, A. M., de Elvira, C., & Sanz, P. D. (2007a). A model to
design high-pressure processes towards an uniform temperature
distribution. Journal of Food Engineering, 78(4), 14631470.
Otero, L., Ousegui, A., Urrutia-Benet, G., de Elvira, C., Havet, M.,
LeBail, A., et al. (2007b). Modelling industrial scale highpressure-low-temperature processes. Journal of Food Engineering,
83(2), 136141.
Ozmutlu, O., Hartmann, C., & Delgado, A. (2006). Momentum and
energy transfer during phase change of water under high
hydrostatic pressure. Innovative Food Science & Emerging
Technologies, 7(3), 161168.
Palou, B., Lopez-Malo, A., & Welti-Chanes, J. (2002). Innovative
fruit preservation methods using high pressure. In J. WeltiChanes, G. V. Barbosa-Canovas, & J. M. Aguilera (Eds.),
Engineering and food for the 21st Century. Food Preservation
Technology Series, Chapter 44 (pp. 314). Boca Raton, FL,
USA: CRC Press LLC.
Patazca, E., Koutchma, T., & Balasubramaniam, V. M. (2007). Quasiadiabatic temperature increase during high pressure processing of
selected foods. Journal of Food Engineering, 80(1), 199205.
Patterson, M., & Kilpatrick, D. (1998). The combined effect of high
hydrostatic pressure and mild heat on inactivation of pathogens
in milk and poultry. Journal of Food Protection, 61(4), 432436.
Pauling, L. (1964). College chemistry: An introductory textbook of
general chemistry. San Francisco, CA: Freeman and Company.
Pehl, M., Werner, F., & Delgado, A. (2000). First visualization of
temperature fields in liquids at high pressure. Experiments in
Fluids, 29, 302304.

Food Bioprocess Technol (2008) 1:234


Perrier-Cornet, J. M., Marchal, P. A., & Gervais, P. (1995). A new
design intended to relate high-pressure treatment to yeast cell
mass transfer. Journal of Biotechnology, 41, 4958.
Polydera, A. C., Stoforos, N. G., & Taoukis, P. S. (2003). Comparitive
shelf life study and vitamin C loss kinetics in pasteurised and
high pressure processed reconstituted orange juice. Journal of
Food Engineering, 60, 2129.
Polydera, A. C., Galanou, E., Stoforos, N. G., & Taoukis, P. S. (2004).
Inactivation kinetics of pectin methylesterase of Greek navel
orange juice as a function of high hydrostatic pressure and
temperature process conditions. Journal of Food Engineering, 62
(3), 291298.
Pontes, L., Cordeiro, Y., Giongo, V., Villas-Boas, M., Barreto, A.,
Araujo, J. R., et al. (2001). Pressure-induced formation of
inactive triple-shelled rotavirus particles is associated with
changes in the spike protein VP4. Journal of Molecular Biology,
307(5), 11711179.
Prstamo, G., & Arroyo, G. (1998). High hydrostatic pressure effect
on vegetable structure. Journal of Food Science, 63, 878881.
Rademacher, B., Werner, F., & Pehl, M. (2002). Effect of the
pressurizing ramp on the inactivation of Listeria innocua
considering thermofluiddynamical processes. Innovative Food
Science & Emerging Technologies, 3, 1324.
Ramirez-Suarez, J., & Morrissey, M. (2006). Effect of high pressure
processing (HPP) on shelf life of albacore tuna (Thunnus
alalunga) minced muscle. Innovative Food Science & Emerging
Technologies, 7(12), 1927.
Rao, M. A., Ooley, H. J., & Vitali, A. A. (1986). Flow properties of concentrated juices at low temperatures. Food Technology, 38, 113119.
Rasanayagam, V., Balasubramaniam, V. M., Ting, E., Sizer, C. E.,
Bush, C., & Anderson, C. (2003). Compression heating of
selected fatty food materials during high-pressure processing.
Journal of Food Science, 68(1), 254259.
Raso, J., & Barbosa-Canovas, G. V. (2003). Non-thermal preservation
of foods using combined processing techniques. Critical Reviews
in Food Science and Nutrition, 43(3), 265285.
Rastogi, N. K., Raghavarao, K. S. M. S., Balasubramaniam, V. M.,
Niranjan, K., & Knorr, D. (2007). Opportunities and challenges
in high pressure processing of foods. Critical Reviews in Food
Science and Nutrition, 47(1), 69112.
Ritz, M., Tholozan, J. L., Federighi, M., & Pilet, M. F. (2002).
Physiological damages of Listeria monocytogenes treated by high
hydrostatic pressure. International Journal of Food Microbiology,
79(12), 4753.
Rodrigo, D., van Loey, A., & Hendrickx, M. E. (2007). Combined
thermal and high pressure colour degradation of tomato puree and
strawberry juice. Journal of Food Engineering, 79(2), 553660.
Roos, Y. H. (2003). Thermal analysis, state transitions and food quality.
Journal of Thermal Analysis and Calorimetry, 71(1), 197203.
Ross, A. I. V., Griffiths, M. W., Mittal, G. S., & Deeth, H. C. (2003).
Combining non-thermal technologies to control foodborne microorganisms. International Journal of Food Microbiology, 89(23),
125138.
Rouille, J., LeBail, A., Ramaswamy, H. S., & Leclerc, L. (2002). High
pressure thawing of fish and shellfish. Journal of Food
Engineering, 53, 8388.
Rubio, B., Martinez, B., Garcia-Cachan, M. D., Rovira, J., & Jaime, I.
(2007). Effect of high pressure preservation on the quality of dry
cured beef Cecina de Leon. Innovative Food Science &
Emerging Technologies, 8(1), 102110.
Sandra, S., Stanford, M. A., & Meunier Goddik, L. (2004). The use of
high-pressure processing in the production of Queso Fresco
cheese. Journal of Food Science, 69(4), 153158.
Sangsuk, O., & Myoung, J. (2003). Inactivation of Bacillus cereus
spores by high hydrostatic pressure at different temperatures.
Journal of Food Protection, 66, 599603.

33
San Martin, M. F., Barbosa-Canovas, G. V., & Swanson, B. G. (2002).
Food processing by high hydrostatic pressure. Critical Reviews in
Food Science and Nutrition, 42(6), 627645.
San Martin-Gonzalez, M. F., Welti-Chanes, J., & Barbosa-Canovas, G.
V. (2006). Cheese manufacture assisted by high pressure. Food
Reviews International, 22(3), 275289.
San Martin-Gonzalez, M. F., Rodriguez, J. J., Gurram, S., Clark, S.,
Swanson, B. G., & Barbosa-Canovas, G. V. (2007). Yield,
composition and rheological characteristics of cheddar cheese
made with high pressure processed milk. LWT-Food Science
Technology, 40(4), 697705.
Saul, A., & Wagner, W. (1989). A fundamental equation for water
covering the range from the melting line to 1273 K at pressures
up to 25000 MPa. Journal of Physical and Chemical Reference
Data, 18(4), 15371564.
Schluter, O. (2003). Impact of high pressure - low temperature
processes on cellular materials related to foods. PhD thesis,
Berlin University of Technology.
Schluter, O., Urrutia-Benet, G. U., Heinz, V., & Knorr, D. (2004).
Metastable states of water and ice during pressure-supported
freezing of potato tissue. Biotechnology Progress, 20(3), 799810.
Schreck, C., Layh-Schmidt, G., & Ludwig, H. (1999). Inactivation of
Mycoplasma pneumoniae by high hydrostatic pressure. Pharmaceutical Industry, 61(8), 759762.
Schubring, R., Meyer, C., Schluter, O., Boguslawski, S., & Knorr, D.
(2003). Impact of high pressure assisted thawing on the quality of
fillets from various fish species. Innovative Food Science &
Emerging Technologies, 4(3), 257267.
Seeton, C. J. (2006). Viscositytemperature correlation for liquids.
Tribology Letters, 22(1), 6778.
Serra, X., Grebol, N., Guardia, M. D., Guerrero, L., Gou, P.,
Masoliver, P., et al. (2007). High pressure applied to frozen
ham at different process stages. 2. Effect on the sensory attributes
and on the colour characteristics of dry-cured ham. Meat Science,
75(1), 2128.
Shen, T., Urrutia-Benet, G., Brul, S., & Knorr, D. (2005). Influence of
high-pressure-low-temperature treatment on the inactivation of
Bacillus subtilis cells. Innovative Food Science & Emerging
Technologies, 6(3), 271278.
Shimada, S., Andou, M., Naito, N., Yamada, N., Osumi, M., &
Hayashi, R. (1993). Effects of hydrostatic pressure on the
ultrastructure and leakage of internal substances in the yeast
Saccharomyces cerevisiae. Applied Microbiology and Biotechnology, 40, 23131.
Slade, L., & Levine, H. (1991). Beyond water activity: Recent
advances based on an alternative approach to the assessment of
food quality and safety. Critical Reviews in Food Science and
Nutrition, (30), 115.
Smelt, J. P. P. M. (1998). Recent advances in the microbiology of high
pressure processing. Trends in Food Science & Technology, 9,
152158.
Smith, A. E., Moxham, K. E., & Middelberg, A. P. J. (1998). On uniquely
determining cell-wall material properties with the compression
experiment. Chemical Engineering Science, 53, 39133922.
Smith, A. E., Zhang, Z., & Thomas, C. R. (2000a). Wall material
properties of yeast cells: Part 1. Cell measurements and
compression experiments. Chemical Engineering Science, 55
(11), 20312041.
Smith, A. E., Moxham, K. E., & Middelberg, A. P. J. (2000b). Wall
material properties of yeast cells. Part II. Analysis. Chemical
Engineering Science, 55(11), 20432053.
Spilimbergo, S., Elvassore, N., & Bertucco, A. (2002). Microbial
inactivation by high-pressure. Journal of Supercritical Fluids, 22
(1), 5563.
Sun, D.-W. (Ed.) (2005). Emerging technologies for food processing.
London, UK: Elsevier.

34
Supavititpatana, T., & Apichartsrangkoon, A. (2007). Combination
effects of ultra-high pressure and temperature on the physical and
thermal properties of ostrich meat sausage (yor). Meat Science,
76(3), 555560.
Taylor, D. M. (1999). Inactivation of prions by physical and chemical
means. Journal of Hospital Infection, 43(Suppl), S69S76.
Thiebaud, M., Dumay, E., & Cheftel, J. C. (2002). Pressure-shift
freezing of o/w emulsions: Influence of fructose and sodium
alginate on undercooling, nucleation, freezing kinetics and ice
crystal size distribution. Food Hydrocolloids, 16, 527545.
Thomas. C. R., & Zhang, Z. (1998). The effect of hydrodynamics on
biological materials. In E. Galindo, & O. T. Ramrez (Eds.),
Advances in bioprocess engineering II (pp. 137170). London:
Kluwer.
Ting, E. Y., & Anderson, C. (2006). Systems and methods to slowly
reduce the pressure in a pressure chamber over time. U.S. Patent
2006272709; (7 Dec 2006).
Ting, E. Y., & Lonneborg, N.-G. (2002). Method and apparatus for
high pressure treatment of substances under controlled temperature conditions. U.S. Patent 20020192109; (19 Dec 2002).
Ting, E. Y., & Marshall, R. G. (2002). Production Issues Related to
UHP Food. In J. Welti-Chanes, G. V. Barbosa-Canovas, & J. M.
Aguilera (Eds.), Engineering and food for the 21st century. Food
preservation technology series, Chapter 44 (pp. 727738). Boca
Raton, FL, USA: CRC Press LLC.
Toepfl, S., Mathys, A., Heinz, V., & Knorr, D. (2006). Potential of
high hydrostatic pressure and pulsed electric fields for energy
efficient and environmentally friendly food processing. Food
Reviews International, 22(4), 405423.
Torres, J. A., & Velazquez, G. (2005). Commercial opportunities and
research challenges in the high pressure processing of foods.
Journal of Food Engineering, 67(12), 95112.
Trejo-Ayara, X. I., Hendrickx, M., Verlinden, B. E., Van Buggenhout,
S., Smale, N. J., Stewart, C., et al. (2007). Understanding texture
changes of high pressure processed fresh carrots: A microstructural and biochemical approach. Journal of Food Engineering, 80(3), 873884.
Upmann, M., Paulsen, P., James, C., & Smulders, F. J. M. (2000).
Microbiology of refrigerated meat. Fleischwirtschaft, 80, 9097.
Urrutia-Benet, G. (2005). High-pressure-low-temperature processing
of foods: Impact of metastable phases of process and quality
parameters. PhD thesis, Berlin University of Technology.
Urrutia-Benet, G., Schlter, O., & Knorr, D. (2004). High pressure-low
temperature processing. Suggested definitions and terminology.
Innovative Food Science & Emerging Technologies, 5(4), 413427.
Urrutia-Benet, G., Chapleau, N., Lille, M., LeBail, A., Autio, K., &
Knorr, D. (2006). Quality related aspects of high pressure low
temperature processed whole potatoes. Innovative Food Science
& Emerging Technologies, 7(12), 3239.
Urrutia-Benet, G., Balogh, T., Schneider, J., & Knorr, D. (2007).
Metastable phases during high-pressure-low-temperature processing of potatoes and their impact on quality-related parameters.
Journal of Food Engineering, 78(2), 375389.
Van Buggenhout, S., Messagie, I., Van Loey, A., & Hendrickx, M. E.
(2005). Influence of low-temperature blanching combined with
high-pressure shift freezing on the texture of frozen carrots.
Journal of Food Science, 70(4), S304S308.
Van Buggenhout, S., Messagie, I., Maes, V., Duvetter, T., Van Loey,
A., & Hendrickx, M. E. (2006a). Minimizing texture loss of
frozen strawberries: Effect of infusion with pectinmethylesterase
and calcium combined with different freezing conditions and

Food Bioprocess Technol (2008) 1:234


effect of subsequent storage/thawing conditions. European Food
Research and Technology, 223(3), 395404.
Van Buggenhout, S., Lille, M., Messagie, I., Van Loey, A., Autio, K.,
& Hendrickx, M. (2006b). Impact of pretreatment and freezing
conditions on the microstructure of frozen carrots: Quantification
and relation to texture loss. European Food Research and
Technology, 222(56), 543553.
Van den Berg, R. W., Hoogland, H., Lelieveld, H. L. M., & van
Schepdael, L. (2001). High pressure equipment for food
processing applications. In M. E. Hendrickx, & D. Knorr
(Eds.), Ultra high pressure treatment of food, Chapter 11 (pp.
297312). London, UK: Kluwer Academia.
Van Loey, A., Ooms, V., Weemaes, C., Van den Broeck, I.,
Ludikhuyze, L., Indrawati, et al. (1998). Thermal and pressuretemperature degradation of chlorophyll in broccoli (Brassica
oleracea L italica) juice: A kinetic study. Journal of Agricultural
and Food Chemistry, 46(12), 52895294.
Van Opstal, I., Bagamboula, C. F., Vanmuysen, S. C. M., Wuytack, E.
Y., & Michiels, C. W. (2004). Inactivation of Bacillus cereus
spores in milk by mild pressure and heat treatments. International Journal of Food Microbiology, 92, 227234.
Villarreal-Alba, E. G., Contreras-Esquivel, J. C., Aguilar-Gonzalez, C.
N., & Reyes-Vega, M. L. (2004). Pectinesterase activity and the
texture of Jalapeno pepper. European Food Research and
Technology, 218(2), 164166.
Vitali, A. A., & Rao, M. A. (1984). Flow properties of low pulp
concentrated orange juice - serum viscosity and effect of pulp
content. Journal of Food Science, 49(3), 876881.
Watson, J. T. R., Basu, R. S., & Sengers, J. V. (1980). An improved
representative equation for the dynamic viscosity of water
substance. Journal of Physical and Chemical Reference Data,
9, 12551290.
Weagent, S. D., Bryant, J. L., & Bark, D. H. (1994). Survival of E.
coli (O157:H7) in mayonnaise based sauces at room and
refrigerated temperatures. Journal of Food Protection, 57, 629
631.
Welti-Chanes, J., Lopez-Malo, A., Palou, E., Bermudez, D., GuerreroBeltran, J. A., & Barbosa-Canovas, G. V. (2005). Fundamentals
and applications of high pressure processing to foods. In G. V.
Barbosa-Canovas, M. S. Tapia, & P. M. Cano (Eds.), Novel food
processing technologies, Chapter 8 (pp. 157181). Boca Raton,
FL, USA: CRC Press LLC.
Wilkinson, N., Kurdziel, A. S., Langton, S., Needs, E., & Cook, N.
(2001). Resistance of poliovirus to inactivation by hydrostatic
pressures. Innovative Food Science & Emerging Technologies, 2,
9598.
Yuste, J., Mor-Mur, M., Capellas, M., & Pla, R. (1999). Pressure vs.
heat-induced bacterial stress in cooked poultry sausages: A
preliminary study. Letters in Applied Microbiology, 294, 233
237.
Yuste, J., Capellas, M., Pla, R., Fung, D. Y. C., & Mor-Mor, M.
(2001). High pressure processing for food safety and preservation: A review. Journal of Rapid Methods and Automation in
Microbiology, 91, 110.
Zhao, Y., Flores, R. A., & Olson, D. (1998). High hydrostatic pressure
effects on rapid thawing of frozen beef. Journal of Food Science,
632, 272275.
Zhu, S. M., LeBail, A., Chapleau, N., Ramaswamy, H. S., & de
Lamballerie-Anton, M. (2004) Pressure shift freezing of pork
muscle: Effect on color, drip loss, texture, and protein stability.
Biotechnology Progress, 203, 939945.

Vous aimerez peut-être aussi