High-Pressure Processing: Fundamentals, Misconceptions, and Advances

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337320568

High-Pressure Processing: Fundamentals, Misconceptions, and Advances

Chapter · January 2019


DOI: 10.1016/B978-0-08-100596-5.22949-1

CITATIONS READS
0 1,232

2 authors:

Kaavya Rathnakumar Sergio I. Martinez-Monteagudo


University of Wisconsin–Madison New Mexico State University
27 PUBLICATIONS   230 CITATIONS    68 PUBLICATIONS   819 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Extraction of phospholipids from dairy byproducts View project

Call for Papers: Our Special Issue "Implications of Drying Techniques on the Bioactive Components and Quality of Food" has launched in Journal of Food Quality. Now
accepting original research and review articles. View project

All content following this page was uploaded by Sergio I. Martinez-Monteagudo on 14 March 2020.

The user has requested enhancement of the downloaded file.


High-Pressure Processing: Fundamentals, Misconceptions, and Advances
Kaavya Rathnakumar, and Sergio I Martínez-Monteagudo, Dairy and Food Science Department, South Dakota State
University, Brookings, SD, United States
© 2019.

Email addresses: Kaavya.Rathnakumar@sdstate.edu (K. Rathnakumar); Sergio.MartinezMonteagudo@sdstate.edu (S.I.

F
Martínez-Monteagudo)
Introduction

OO
High-pressure refers to a hydraulic power system that uses pressure levels in the range of 100 to 600 MPa. In such a hydraulic system, a
food material (liquid or solid) is hydrostatically compressed by means of a fluid, which produces nearly uniform change in the molecular
volume, and consequently desirable modifications in the material. Examples of such modifications are the crystallization of fat, phase
transition, ionization, inactivation of enzymes, denaturation of proteins, inactivation of microorganisms, and delaying the rate of some
chemical reactions.
In a short span of approximately 30 years, high-pressure has transitioned from a proof-concept carried out in few research laboratories

PR
to an established unit operation, adopted in multiple food plants. Nowadays, high-pressure products including deli meats, seafood, fruit
juices, vegetable products, salads, and salsa are readily available in the supermarkets of North America, Europe, and Asia-Pacific with a
volume of sales estimated at over $2 billion annually.
Around the globe, scientists and engineers have performed basic and applied research in universities, national laboratories, and
industries. These research activities have resulted in significant advancement in different aspects of the technology, including fundamental
compression studies, physical properties, process uniformity, and structural modifications. These technological developments have led to
a multitude of applications, from pasteurization to extraction. The significant growth in the research activities of high-pressure has led

D
to a vast amount of literature on the subject, and it becomes increasingly incomprehensible to review the state-of-the-art. The notorious
developments of high-pressure also bring some misconceptions as well. This chapter covers an overview of the historical progress,
fundamentals of high-pressure, equipment overview, some common misconceptions, and a summary of recent advances.
TE
Historical Progress

In this historical overview, those milestones that the authors considered as the forerunner in the development of high-pressure are included.
EC

Many early experiments, particularly those dealing with the compressibility of liquids (1760s), devising accurate methods for measuring
pressure (1850s), and melting of solids (1890s) are not included. The industrial adaptation of high-pressure applications is a result of many
technological milestones over centuries. More specifically, the foundations, development, and implementation of high-pressure has been
tight to two major disciplines: mechanical and chemical engineering. Table 1 lists some remarkable milestones that directly and indirectly
helped the development of high-pressure processing as currently known. The invention of the steam engine in 1769 by James Watt was
perhaps the first relevant milestone regarding the use of high-pressure. Watt's invention allowed to transfer thermal energy into kinetic
RR

energy at elevated pressures (Eggers, 2005). By 1826, it was observed that the boiling point of water raised with increasing the pressure.
Years after, these observations led to the discovery of the critical point of fluids including water and CO2 (Dewar, 1882), which in turns
laid the foundation of supercritical fluid technology. By mid-19th century (1820–60), industrial operations at pressure levels of tens MPa
had become a reality. By 1884, the American Society of Mechanical Engineers published the first standard test methods of boilers operated
at high pressures.
CO

The ability of pressure to influence the course of chemical reactions was a topic of research interest in the late part of the 19th century
(1850–99), where the foundations of chemical kinetics in the gas phase at elevated pressures enabled the hydrogenation process, which
later became an essential step during the synthesis of fine chemicals. In 1881, the hydrolysis of starch to produce glucose monomers was
performed under pressure (Soxhlet, 1881). Years later, it was observed that the application of 50 MPa inhibited the acid inversion of sucrose
(Stern, 1897). Such early milestones laid the foundations for high-pressure chemical synthesis. A remarkable example is the synthesis
of ammonia from nitrogen and hydrogen at 20 MPa by Haber and Bosch (Nobel prize in 1918) (Eggers, 2005). Soon after, many other
chemical syntheses were developed into commercial processes including methanol, phenols (Brown, 1920), polyols, and polymerization
UN

of ethylene using pressure levels of 300 MPa (Vetter, 2001). The development of seamless and forged cylindrical components enabled
the safe operation and control of the vessels at pressure levels in the range of hundreds MPa. The production of synthetic diamonds from
graphite at pressures above 1200 MPa became a reality in the mid-20th century.
Behind such landmarks is the development of pumps, compressors, tubes, seals, and sensors. Altogether, it opened a window to an
unexplored region of high pressure. Sheet metal forming and isostatic pressing of advanced materials are notable examples of industrial
applications of high pressure.
Operating at 680 MPa, Hite (1899) prevented the spoilage of milk by killing some microorganisms. Hite's experiment was the first
application of high pressure aimed at extending the shelf-life of foods. In 1900, Auguste Gaulin invented the first apparatus for applying
high-pressure into fluid milk, reaching pressure levels 15 MPa (Pandolfe, 1982). Gaulin's invention aimed at reducing the size of the
milk fat globule that led to the development of a critical unit operation, homogenization, of dairy manufacturing. Between the early and
mid-20th century (1909–59), Percy W. Bridgman, who is considered the father of high-pressure, performed fundamental studies dealing

Innovative Food Processing Technologies: A Comprehensive Review, Volume ■ https://doi.org/10.1016/B978-0-12-815781-7.22949-1 1


2 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

Table 1 Selected developments in the industrial application of high pressure

Year Milestone Reference

1769 Invention of the steam engine Eggers (2005)


1826 Experiments on boiling point of water Vetter (2001)
1860 Discovery of the critical point of fluids Dewar (1882)
1881 Conversion of starch into sugar Soxhlet (1881)

F
1897 Inversion of cane sugar Stern (1897)
1899 Experiments on milk preservation Hite (1899)
1900 Homogenization of milk Pandolfe (1982)

OO
1909 Experiments on compressibility Bridgman (1909)
1914 Water phase diagram Bridgman (1914a)
1914 Coagulation of albumen Bridgman (1914b)
1918 Pressure effect on bacteria Larson et al. (1918)
1920 First commercial plant of ammonia Vetter (2001)
1920 Phenol synthesis Brown (1920)
1923 Experiments on thermal conductivity Bridgman (1923)

PR
1943 Mutarotation of glucose Sander (1943)
1953 Synthetic diamonds Vetter (2001)
1978 Commercial decaffeination plant King (2014)
1980 Beef protein quality Kennick et al. (1980)
1990 Commercial application of food products Mozhaev et al. (1994)
1995 Experiments on pressure-assisted freezing Kalichevsky et al. (1995)
1997 Introduction of pressure treated guacamole Sizer et al. (2002)

D
2002 Early observation of pressure-assisted extraction Wanehaya et al. (2002)
2004 Pressure as a pasteurization method Balasubramaniam et al. (2015)
2009 FDA issued no objection to an industry petition Balasubramaniam et al. (2015)
2014 Development of pressure-ohmic-thermal sterilization Park et al. (2014)
TE
2015 FDA approved the process pressure enhanced sterilization IFSH (2015)

Adapted from Balasubramanian et al. (2015).


EC

with the compressibility of fluids (Bridgman, 1909), thermal conductivity (Bridgman, 1923), and phase change (Bridgman, 1914a).
Bridgman elucidated the phase diagram of water over a wide range of pressure and temperature. For the first time, a phenomenological
description of the physical state of water was provided. This is a significant milestone that greatly benefited the food industry since several
processes have been designed and optimized according to the phase diagram of water. Another relevant milestone in the field of food
science was the coagulation of albumen by pressure, whose appearance resembled that of the hard-boiled egg (Bridgman, 1914b). This
demonstration was an early indicator of the potential use of high-pressure for structural modifications of food proteins. Another relevant
RR

early observation was reported by Sander (1943), who significantly increased the speed of mutarotation of glucose with the application of
high-pressure. Soon after, it was discovered that pressure killed non-spore-forming microorganisms, while bacterial spores resisted pressure
levels up to 1200 MPa (Larson et al., 1918).
In 1978, the decaffeination process using supercritical carbon dioxide was commercially established. By the 1980s, a considerable
number of research papers were available documenting the effects of high-pressure on microbes, starches, and various proteins including
CO

enzymes. Kennick et al. (1980) demonstrated that pressurization of pre-rigor muscles produces very tender form meat. Despite all the
research, the operational and maintenance costs were the main limiting factor for the commercialization of food products treated by
high-pressure. Nevertheless, a Japanese company in the early 1990's launched the first pressure treated product, jams and jellies. Such
a milestone marked the beginning of a new era for the food industry (Mozhaev et al., 1994). In 1995, the potential application of
high-pressure for freezing and thawing was documented (Kalichevsky et al., 1995). Soon after (1997), Fresherized Foods (formerly
Avomex) launched the first pressurized food product in the USA, an avocado paste made from fresh ripe avocados (Sizer et al., 2002).
In 2002, a novel technique for the digestion of metals in solid matrices was developed using a programmed sequence of temperature and
UN

pressure (Wanekaya et al., 2002), which demonstrates the potential application of high-pressure for enhancing the extraction in food
materials. By 2004, the National Advisory Committee on Microbiological Criteria for Foods recommended to the federal regulators to
redefine the term of pasteurization. The agency redefines the term as any process treatment (such as high-pressure) or combination thereof,
that is applied to food to reduce the most resistant microorganism(s) of public health significance to a level that is not likely to present a
public health risk under normal conditions of distribution and storage. In 2009, the Food Drug Administration agency (FDA) approved the
application of high pressure to a preheated sample for sterilization of low-acid foods (Balasubramaniam et al., 2015). From a regulatory
point of view, this is a significant milestone for the commercialization of sterile foods, which is expected to be on the short to medium term.
In the following years, pressure pasteurized products including deli meats, seafood, fruit juices, vegetable products, salads, and salsa are
readily available in the supermarkets in North America, (USA, Canada), Europe and Asia-Pacific with a volume of sales estimated at over
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 3

$2 billion annually. More recently, a new technology was developed named pressure-ohmic-thermal sterilization to reduce thermal damage
in vegetables (Park et al., 2014). In 2015, the FDA approved the process of pressure enhanced sterilization for the commercial production
of complex particulate-bearing foods.

Governing Principles

F
As discussed earlier, the principles of high-pressure are based on mechanical and chemical engineering. In this section, such governing
principles are presented and discussed.

OO
Definition of Pressure
Pressure can be defined as the ratio of a force divided by the area which that force acts on (Skelton and Webb, 2003). This definition is
rather vague since the numerical description of a solid subjected to a force (pressure) not only depends on the direction in which it acts but
also on the orientation of the surface upon which it is acing. These considerations are decomposed into nine components, stress tensors,
and they are derived from a Cartesian coordinate system (x, y, and z coordinate directions). The nine components represent the direct stress

PR
(compressive) and the shear stress (tangential) induced by the application of pressure.
The units of pressure are force per unit of area, Newtons per square meter = 1 Pa or pounds per square inch. A number of unit operations
in the food industry involve the application of pressure, ranging from ten to hundred MPa (Table 2). Extrusion is a thermal process that
may employ modest pressure. It is a process in which the ingredients are discharged through an opening in a perforated plate or die.
Spray-drying involves the application of pressure for the atomization of particles. Combinations of pressure and temperature yielding water
in the subcritical state, i.e., water above the boiling and below the critical point, have been used for the extraction of bioactive compounds.
Similarly, supercritical pressures have been used in decaffeination and similar food processing operations.

D
Hydrostatic Condition
TE
Pressurization of a material employing a fluid is the essential element of hydrostatic pressure. The hydrostatic term refers to the physics of
contact forces acting on fluids at rest, fluids without any global motion, and it is founded on two principles, equilibrium and transmission
of the force. Microscopically, the pressure within a fluid is a result of direct contact between molecules, where the average molecular
transfer of momentum on a given surface produces the acting force known as pressure. In a particle or a body immersed in the fluid,
the acting force is equally distributed in all directions, a phenomenon known as hydrostatic equilibrium. This is because the pressure
EC

is the only contact force that is possible in a fluid at rest, and it acts on the surface of a body. The second principle states that all
fluids yield to a force imposed on it, transporting the force among themselves without friction and acting without friction on subsequent
contact surface. Under a hydrostatic condition, the diagonal components of the stress tensor are the same value, meaning that the resulting
shear component is zero. Thus, a true hydrostatic condition will not subject a test specimen directly to shear stress, and it is considered
that the hydrostatic condition is independent of time, space and magnitude (Martínez-Monteagudo and Balasubramaniam, 2016). In
high-pressure applications where a packed food material is surrounded by pressuring fluid, the pressure and, more importantly, its effects
RR

are often considered to be instantaneous and homogenous within the food item, regardless of geometry and size. This unique characteristic
has enabled processes development, contributing to successful commercial applications (Balasubramaniam et al., 2015).

Atomic Changes
The most noticeable effect induced by hydrostatic pressure is the uniform change in volume of a fluid or solid, whose magnitude is greater
CO

than the volume changes induced by temperature. A system composed of molecules and atoms forms nuclei of surrounding electromagnetic
forces, and on which an external hydrostatic pressure reduces the internal potential energy (PE) while increasing the internal kinetic (KE)
energy according to Eqs. (1) and (2) (Bridgman, 1935
UN

Table 2 Examples of industrial operations of high-pressure in food industry

Unit operation Pressure (Pa)

Extrusion 1.0 × 106


Spray drying 2.0 × 107
Supercritical fluid technology 3.0 × 107
High-pressure homogenization 2.0 × 108
High-pressure processing 6.0 × 108
4 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

).

(1)

(2)

F
Where E is the total energy of the system, p is the hydrostatic pressure, and v is the volume. In an idealized system, the increase
in the internal kinetic energy due to the hydrostatic pressure is a direct result of the increment of the electron volts per atom, which
results in a shrinkage of the orbits, a compression of the atom. Such theoretical interpretation has been exemplified by the ionic

OO
lattice of the NaCl (Bridgman, 1935). The ionic lattice is held together by the electrostatic attractions of the ions and prevented from
collapsing by a repulsive force due to ionic interpenetration. The magnitude of the attractive forces can be derived from the lattice
structure and the ionic charges, while the component of repulsive force is somewhat related to the distance between atoms. Under
this consideration, the repulsive force obeys the law of force between atoms, where the assemblage of the entire lattice is a function
only of the distance between atomic centers irrespective of the orientation (orientation is negligible for hydrostatic pressure). According
to the law of force approximation, hydrostatic pressure alters the interatomic distance affecting only those interaction whose bonding

PR
energy is distance-dependent (Martinez-Monteagudo and Saldaña, 2014). For instance, the force of electrostatic interactions is inversely
proportional to the distance between charged particles, whereby the application of pressure will affect its bonding strength. Hydrogen
bonding and van der Waals forces are also distance-dependent and therefore are greatly affected by pressure. However, covalent bonds are
unlikely to be affected by pressure because their bonding distance can be minimally compressed further. Indeed, studies have revealed that
covalent bonds that constitute the primary structure of proteins are unaffected by pressure (up to 1500 MPa) (Mozhaev et al., 1994). Table
3 exemplifies the effect of hydrostatic pressure on chemical bonds based on their interatomic distance. The ability of hydrostatic pressure
to keep covalent bonds unaffected has been the central hypothesis for the preservation of biological activity of functional compounds, such

D
as ascorbic acid (Indrawati et al., 2004), folates (Butz et al., 2004), antioxidants (Matser et al., 2004), anthocyanins (Verbeyst et al.,
2010), lycopene (Yan et al., 2017), and conjugated linoleic acid (Martínez-Monteagudo et al., 2012).
TE
Pressure-Volume-Temperature Relations
The assumption of the law of force provides a qualitative approximation of the shrinkage of the ionic lattice and applied pressure,
compressibility. Over the years, this approximation has been redefined in terms of the estimation of the internal kinetic energy, and it has
brought a thermodynamic relation between the deformation of the atoms and the change of internal energy (E) with pressure, Eq. (3).
EC

(3)

Thermodynamically, a substance is completely characterized when the relation of pressure-volume-temperature is known. Eq. (3)
explains the molecular changes induced by pressure. Examples of such changes are density, viscosity, solubility, compressibility, melting,
RR

as well as electrical, thermophysical, and optical properties.


A closer examination of the Eq. (3) reveals that the application of pressure decreases the internal energy (the term of the left side
of Eq. 3), meaning that more energy flowing out in the form heat to counteract the mechanical energy work produced by the external
pressure. This phenomenon is known as adiabatic heat or heat of compression. Most of the foods containing high moisture produce
values of heat of compression similar to that of water, 3 °C per 100 MPa. On the other hand, the heat of compressing foods rich in
CO

fat is on average 8.7–9.1 °C per 100 MPa. Table 4 reports values of heat of compression for selected foods. The numerical value of
the heat of compression is of technological interest for enhancing the thermal processing of a preheated sample, a technology known
as pressure-assisted thermal processing (PATP) or Pressure Enhanced Sterilization (PES). The idea behind this technology is to use the
heat of compression to reach commercial sterilization temperatures (120 °C). Samples compressed hydrostatically rise their temperature
pseudo-instantaneously and uniformly throughout the sample. Such rapid heating reduces the severity of traditional thermal process and the
UN

Table 3 Illustration of the influence of hydrostatic pressure on selected chemical bonds

Type of interaction Working distance [nm] Distance dependence Possible pressure effect

Coulomb 20 Proportional inversely Highly affected


van der Waals 1–20 Optimum distance Highly affected
Hydrogen bonding 0.2 Quadratic Affected
Solvation >2 Exponential Affected
Hydrophobic >2 Exponential Affected
Covalent 0.2 Independent Unlikely affected

Reprinted by permission from Springer Nature: Food Engineering Reviews Martinez-Monteagudo and Saldaña (2014).
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 5

Table 4 Values of heat compression for selected food products

Food Heat of compression (°C per 100 MPa)

Water 3.0
Milk 3.0
Yogurt 3.1
Honey 3.2

F
Salmon 3.2
Olive oil 8.7
Soy oil 9.1

OO
Adapted from Balasubramanian et al. (2015).

lack of temperature uniformity that occurs in a traditional thermal process (Sevenich and Mathys, 2018). PATP was first developed for
the sterilization of low-acid foods (Sizer et al., 2002).

PR
Le Chatelier's Principle
The Le Chatelier principle has a basis in the first and second law of thermodynamics. For a reversible process, the changes in the internal
energy are the product of the relations temperature-entropy and pressure-volume, Eq. (4):

(4)

D
In general terms, the Le Chatelier principle stated that a system in equilibrium would shift to new equilibrium to partially undo an
external force, action, or stress that disturbs the equilibrium. The Le Chatelier's principle is often used to explain the effects of pressure and
TE
temperature on chemical, biological, and physical phenomena. Despite its various applications, the Le Chatelier's principle is a very general
statement about systems in equilibrium and their behavior when subjected to external force or stress. In 1920s, the term of affinity was
developed based on the thermodynamic conjugate pairs, extensive and intensive variables. Thermodynamic variables come in conjugate
pairs: temperature and entropy; pressure and volume; chemical potential and moles. An interpretation of affinity is that such system held at
fixed entropy and volume will come to equilibrium by varying temperature and pressure. In the case of HPP, if pressure (extensive variable)
EC

changes, the equilibrium shifts in the direction that tends to reduce the change in the corresponding intensive variable (volume). Thus, any
phenomenon (phase transition, change in molecular configuration, chemical reaction) accompanied by a decrease in volume is enhanced
by pressure.

Transition State Theory


RR

The Transition State Theory (TST) was developed by Eyring and Polanyi (Eyring, 1935), and it stated that an elementary chemical reaction
proceeds when the molecular positions of their reagents rearrange into a critical position or activated state (Murzin and Salmi, 2005). Fig.
1 illustrates the transition state and activated complex of a hypothetical reaction in solution. Essentially, the TST considers thermodynamic
and statistical mechanics principles to predict the number of combinations of reactants that will be present in the transition state. It is
assumed that interphase exists between the region of reactants and products. This dividing interphase is located at the transition state,
maximum value of the potential energy surface on the minimum energetic path that connects reactants and products. This state is induced
CO

by the collision of molecules, and it the most energetically favourable position. Any trajectory passing through the dividing surface from
the reactant side is assumed to eventually form products. During the transition, the reactants and activated complexes are assumed to be
in some type of equilibrium maintaining their Boltzmann energy distributions corresponding to the temperature of the reacting system,
while the whole system is not at equilibrium. The activated complex has the same properties of an ordinary molecule, meaning that
Maxwell-Boltzmann distribution governs their energy profile within the transition state (Murzin and Salmi, 2005). Fig. 2 exemplifies the
Maxwell-Boltzmann distribution. As the temperature increases, the population of molecules with more energy also increases. From the
UN

Maxwell-Boltzmann distribution, it can be evident that higher temperatures produce not only higher average kinetic energy but also broader
energy range. Once the products are formed, their potential energy decreases and the number of molecules in the transition state is small
compared with the number of reactants. The probability of finding a particle in the transition state depends on the Maxwell-Boltzmann
energy distribution and the partition function. According to statistical physics, the partition function for a particle is a result of the sum of
energies derived from electronic, vibrational, rotational, and translational. Since the activated complex is a point on a potential surface, the
rate of a reaction given by the equilibrium of activated complexes near the top of the potential barrier multiplied by the average velocity of
crossing the barrier. Mathematically, the rate constant (ki) is given by Eq. (5)
6 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

F
OO
PR
Figure 1 Energy profile of a hypothetical reaction illustrating the transition state theory. Reprinted from Catalytic Kinetics, Murzin and Salmi
(2005), Chapter 3 - Elementary reactions with permission for Elsevier.

D
TE
EC
RR

Figure 2 Representation of Maxwell-Boltzmann distribution for a hypothetical reaction. Reprinted from Catalytic Kinetics, Murzin and Salmi
(2005), Chapter 3 - Elementary reactions with permission for Elsevier.
CO

(5)
UN

where k and h are the Boltzmann and Planck constants, respectively, and ΔG≠ is the standard free energy change between reagents and
transition state. The last term of Eq. (5) is the activity coefficient. Partial differentiation of Eq. (5) yields Eq. (6), which introduces the term
activation volume (ΔV≠):

(6)
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 7

The term ΔV≠ is the difference in partial molar volumes between reagents and transition state. Eq. (6) predicts an increase in the rate
when the transition state occupies a smaller volume than the volume of reagents, negative values of activation volume. The activation
volume is perhaps the most important parameter derived from kinetic studies. The sensitivity of a chemical reaction to pressure depends
on the absolute value of ΔV≠ (positive or negative). Reactions with large positive ΔV≠ values may occur at a very low rate with increasing
pressure. Contrary, reaction with negative values of ΔV≠ are favored by the application of pressure. Although the transition state theory
and its associated activation volume was originally developed for chemical reactions, it has been widely used to explain the effect of
pressure in many applications such as fat crystallization, starch gelation, enzyme inactivation, protein denaturation, and inactivation of

F
microorganisms.

OO
Equipment Overview

The methods for the generation of high-pressure are expensive to build, operate, and maintain. Thus, the application of high-pressure will
only be successful if clear advantages are obtained with respect to the product quality, mitigation of food process contaminants, and overall
costs (Sevenich et al., 2015). This section provides an overview of the basic components and their main function of high-pressure units.

PR
Hydraulic Configuration
In a nutshell, HPP units are essentially hydraulic power systems, where the power generated by a compressed fluid is delivered from its
point of generation to its place of deployment. This energy transfer is carried out in different arrangements. The scope of this section is to
introduce the fundamental principles and the basic configurations of HPP and major composing elements.
An HPP unit uses a pump, an intensifier, a control valve, a hydraulic actuator, a fluid reservoir, at least a vessel, and connecting pipes
(Fig. 3). One of the most noteworthy features of HPP units is its power-carrying media or pressure-transmitting fluid, which allows the

D
hydraulic potential energy to be delivered to all directions with the same capacity. This feature allows the arrangement of multiple vessels
being equally powered by the same pump-intensifier. The schematics showed in Fig. 3 are a simplified view of how the components are
connected and their basic function. Actual structure, installation, capacity, and parameters of the individual components will depend on the
TE
specific unit and equipment vendor.

Pressure Generation
In HPP, the pressures commonly used (100–600 MPa) cannot be achieved by a pump directly. The target pressure is thus obtained by
EC

the combined action of a hydraulic pump and a pressure intensifier. A hydraulic intensifier is a mechanical device for intensifying the
hydraulic power at low pressure into a reduced volume of higher pressure. The underlying physics of an intensifier is based on the
Pascal's law for incompressible fluids. Overall, the generation of pressure and its multiplication is conducted by the combined action of
a hydraulic pump and an intensifier. The hydraulic pump not only converts mechanical kinetic energy into hydraulic potential energy but
RR
CO
UN

Figure 3 Simplified diagram of the principal components of high-pressure processing. (1) fluid reservoir, (2) hydraulic pump, (3) intensifier, (4)
check valve, (5) needle valve, (6) decompression valve, (7) pressure transducer, and (8) high-pressure vessel.
8 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

also provides the intake flow of the intensifier, which operates on the ratio-of-areas principle. As illustrated in Fig. 4, a typical intensifier
(single-acting intensifier) has a common rod connecting two pistons of different sizes: 1) the actuator piston, larger piston; and 2) the
boosting piston, smaller piston. The large cylinder has a port connecting the flow corresponding from the hydraulic pump, while the small
cylinder has a port connecting to the low-pressure fluid supply lines, and the high-pressure outline line. A set of check valves separates
the low- and the high-pressure lines. During the pressure generation, the low-pressure fluid is supplied to the large cylinder and acts on
the larger piston driven the smaller piston to discharge pressurized fluid. The ratio between the large and the small cylinder is inversely
proportional to the areas ratio, according to Eq. (7):

F
(7)

OO
where p is the pressure generated by the hydraulic pump, low-pressure; P is the final pressure; R and r are the radio of the large and small
cylinder; and F is a constant related the friction. In general, an intensifier is designed to magnify between 6–10 times the input pressure
value. The actuating cylinder is normally retracted by the fluid. The pressure intensifier illustrated in Fig. 4 boosts pressure intermittently
– each piston stroke discharges pressurized fluid. Single acting-intensifier can be grouped in a pair to provide more continues supply of
pressure. Another configuration of the intensifier is the double-acting intensifier, illustrated in Fig. 5. Essentially, a double-acting intensifier

PR
is a combined design of two single acting-intensifier sharing a large cylinder. The two boosting piston and the actuating piston are mounted
on one rod. This design allows the retraction of one of the boosting pistons, while the other is extending with every stroke of the actuating
piston. The retraction of a boosting piston drawn fluid from the inlet line into the compression chamber. Reciprocally, the extension of a
boosting piston discharges the fluid into the high-pressure outlet. Several commercially available intensifiers are on the market actuated by
a hydraulic pump, which makes it unnecessary for the in-house construction.

D
Transmitting Fluid
The power-carrying or transmitting fluid performs various functions. The primary function is the transmission of power or pressure. For
this task, the fluid should be little compressible over the entire operating range of pressure. All hydraulic fluids experience some degree
TE
of compressibility in proportion to the operating pressure. For instance, water presents compressibility values of about 15% at 600 MPa
(Mathys and Knorr, 2009). Although extreme changes in the compressibility cause delay in the response rate, the compressibility of fluids
commonly used in HPP is not a concern. On the other hand, the presence of dissolved air or gases negatively impacts the response and
performance of the fluid. Thus, one critical property of the transmitting fluid is the percentage of gases that can be dissolved in the fluid.
A secondary function of the transmitting fluid is lubrication, which minimizes wear and reduces friction. Overall, the transmitting fluid
EC

provides sufficient lubricity by forming a film that separates metal-to-metal contact between surfaces. In case of insufficient lubricity,
antiwear additives are used for improving the lubricity capability. A third function of the fluid is to remove contaminants between gaps and
crevices. Contaminants are small particles from surfaces that torn off, and they can accelerate wear.
In order to perform such key functions, the transmitting fluids are evaluated based on their properties including specific gravity,
viscosity index, bulk modulus, antifoaming, lubricity, and oxidation stability. Another important parameter of the transmitting fluids is
RR
CO
UN

Figure 4 Illustration of a single-acting intensifier.


High-Pressure Processing: Fundamentals, Misconceptions, and Advances 9

F
OO
PR
Figure 5 Illustration of a double-acting intensifier.

D
the bulk modulus, which is a measure of the compressibility. The basic definition of the fluid bulk modulus is the fractional reduction in
TE
the fluid volume with respect to the applied pressure, Eq. (8).

(8)
EC

where β is the bulk modulus, V is the volume, and p is the pressure. The bulk modulus of a fluid can be either defined as the isothermal
tangent bulk modulus if its compressibility is measured under a constant temperature, or as the isoentropic tangent bulk modules if its
compressibility is measured under constant entropy.
Another relevant property of transmitting fluid is the antifoaming ability. The formation of foam contains entrained air, which reduces
the dynamic responses in power transmission due to decreased bulk modules, shortening fluid service life because of increased fluid
oxidation. Under certain conditions, hydraulic fluids may dissolve a certain amount of air. In general, the dissolving rate of gases in fluids is
RR

proportional to the pressure and inversely proportional to the temperature. During pressurization and depressurization cycles, the dissolved
gases are released from the fluids forming foams. Oxidation and thermal stability are other important properties of the transmitting fluids.
The resistance to oxidation is known as oxidative stability, and the oxidation rate increases as the temperature rises. Thermal stability is the
capability of the fluids in resisting chemical reactions and further decomposition during high-temperature operations. The seal compatibility
of fluid is one more property that needs to be considered for high-pressure applications. The fluid selected should be compatible with
CO

the sealing materials. While water is commonly used as transmitting fluid in HPP, other transmitting fluids such as petroleum-based and
environmentally safe fluids are used for PATP applications. Synthetic esters, polyglycols, vegetable oils, and water-glycol mixtures are
examples of transmitting fluids used in PATP.

Pressure Vessels
It serves as a point of contact between the packed food and transmitting fluid. Pressure vessels that are intended for use in HPP require
UN

specific design features, the shape of the vessel and thickness of the wall. There are three types of high-pressure vessels used for HPP: i)
thick-walled or monobloc, ii) multiwalled, and iii) wire-wound (Balasubramaniam et al., 2016). The first type of vessel is essentially a
cylindrical wall consisting of a single layer (nonwelded), avoiding large side ports in the wall, with the end closures being for openings
and connections. Monobloc vessels are suitable for a wide range of pressure levels and application that require high temperatures (PATP).
The associated thermal stresses from heating and cooling are rather low because of the thermal conductivity across the wall. The monobloc
vessels are particularly suitable for laboratory units, where heating and cooling can be accomplished quite reliable. Another feature of
the monobloc vessels is the large wall thickness that depends on the operational pressure. Due to the nature of hydrostatic pressure,
tangential stress is highest at the inner bore and decays toward the outer diameter. The monobloc vessels are designed according to
the maximum allowable pressure at a given ratio of the outer diameter to inner diameter, and yield strength of the material (Vetter,
2001). The second type of vessels, multiwalled, is suitable for operating at larger volumes and higher-pressure levels than that of
10 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

the monobloc vessel. However, the thermal conductivity through the multiwall is lower than the monobloc, an aspect that should not be
overlooked for application that involves extensive heating and cooling. Multiwalled vessels are manufactured by autofrettage technique,
where the inner part of a cylindrical vessel is compressed until experiences plastic deformation, while the outer part only experiences
elastic deformation. Upon decompression, the outer part of the cylinder tends to regain its original form, and the net results is a constant
condition of pre-stressed between the plastic and elastic deformation. Another technique for the manufacturing of multiwalled vessels is the
heat-shrink technique, where two separate cylinders, outer and inner, are assembled by heating the outer cylinder to induce its expansion,
while the inner cylinder is simultaneously cooled to induce shrinking. During the shrinking process, compressive stresses are produced in

F
the smaller cylinder, and tensile stresses in the larger cylinder, both in the circumferential and longitudinal directions. The third type of
vessel is the wire-wound, which is a variation of the multiwalled vessel. The manufacture of wire-wound vessels combines a number of
turns of high tensile strength wire wound around an inner cylinder, which shrinks the inner diameter of the cylinder. The shrinking of the

OO
inner cylinder is a result of the pressure derived from the wire, leading to compressive and residual stresses inside the vessel and providing
great strength.

Closures
The type of closure is an important consideration for the HPP vessels. Generally, closures can be either attached to the vessel or supported

PR
by an external frame. In the first case, the vessel not only experiences circumferential and radial stresses but also axial load from the
attached closure. Contrary, the use of an external frame minimizes the axial load reduces the fatigue of the vessel. However, an external
frame results in larger and heavier equipment. The end closures are designed with the idea of reducing cyclic stress and fatigue as well
as facilitating the opening and closing. An end closure with continuous internal thread benefits the fatigue and cyclic stress of the vessel
but it prolongs the time needed for opening and closing. Another design is the end closure connected by bolts. The bolts can be tightened
hydraulically or pneumatically. Despite that a bolted closure reduces the cyclic stress, its operation is very time consuming since each bolt
is manipulated individually. A clamp supported by an external frame may be sued in conjunction with an interrupted thread or locked-type

D
system. This arrangement can be automatized, simplifying the overall operation.

Seals
TE
Leaking is one of the most common occurrences in HPP units, which negatively impacts the hydraulic performance of the entire unit.
Proper sealing prevents leakage without adding much friction under the entire operational conditions. Another function of the sealing is to
exclude contaminants that shorten the service life of the unit. In high-pressure applications, there are two types of seals, static and dynamic
seals. The former is used to seal the clearance between immobile components, and the latter is for applications with relative motion between
EC

surfaces. The static seals act within the clearance of two surfaces to stop fluids from leaking from the gap while allowing the components
to move freely. Factors such as the size of the gap, the length of the seal, the applied pressure, and the smoothness of the surfaces determine
the sealing effectiveness. Static seals act on small gaps which therefore results in low friction and low generation of heat. Metallic seals,
ring joints, conical-, flat-, lapped-, and O-rings are examples of static seals used in HPP units.
On the other hand, dynamic seals should match the microscopic irregularities of the contacting surfaces to prevent pressure fluid
penetration. A proper dynamic seal should have enough flexibility to expand or compress to fit the gap variation caused by the mechanical
RR

movement. In addition, dynamic seals must withstand shear stress produced by the system and to resist being forced into the extrusion gap.
The fundamental principle of sealing for HPP is that the pressure itself provides the sealing element. A suitable seal is normally selected in
terms of pressure, transmitting fluid, and temperature. Elastomer O-ring is frequently used to seal high-pressure connections. O-ring seals
have a round shape cross-section and they are made of oil-resistant rubber. Under pressure, the O-ring is pushed by the high-pressure fluid
to the low-pressure side of the groove, resulting in a tighter seal. These seals are simple, robust, and effective. Cyclic stress creates fatigue
problems in the seal. Seals are not only used for connecting pipes but also for the junction of two moving parts (piston and cylinder).
CO

Fluid Reservoir
The fluid reservoir is a storage tank for holding the transmitting fluid. The reservoir may vary in size, shape (rectangular, cylindrical,
T-shaped, or L-shaped), and material of construction (steel, stainless steel, aluminum, or plastic). Reservoirs are designed to store between
two to four times the pump flow. This consideration helps to maintain a continuous supply of fluid in the intensifier and the unit as a
UN

whole. Intensifiers having cylinders with large rods require a constant supply of fluid since the volume of the returning fluid is greatly
reduced with every stroke of the piston. Moreover, a large reservoir provides sufficient cooling capability since the heat of the returning
fluid is dissipated within the excess of fluid and subsequently within the exterior walls of the reservoir. A well-designed reservoir
performs many other functions such as facilitating the settling down of contaminants, releasing of entrained air, preventing the return
of fluids, removing used fluids and contaminants, and indicating fluid levels for maintenance. In a way, a fluid reservoir is much more
than merely a fluid storage tank. Reservoirs for HPP are equipped with an assembly of supporting elements (Fig. 6), including a filter
cap, a breather cap, a fluid level indicator, a separation baffle, an outlet filter, a fluid returning pipe, and a drain plug. The filler cap is
designed to add new fluid, and it is normally located at the top of the reservoir. On the other hand, the breather cap is used to prevent a
vacuum inside the reservoir that will stop the fluid from flowing out the reservoir. A common feature of a reservoir is a baffling device
which prevents the returning fluid from being directly recirculated back into the system without mixing or allowing to settle down or
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 11

F
OO
PR
D
TE
Figure 6 Illustration of a typical fluid reservoir. (1) returning fluid, (2) baffle, (3) filter cap, (4) breather cap, (5) outlet with filter, (6) level
indicator, and (7) drain plug.
EC

discharge air. A magnetic drain plug is perhaps the most effective way to keep metal-based contaminant near the plug and prevent those
contaminants from being carried away into the intensifier. A simple and reliable way to detect malfunction within the system is by the
routine inspection of the fluid level indicator, which is attached on the outside wall of the reservoir. The outlet filter is used to prevent large
solids contaminant s from entering the intensifier.
RR

Misconceptions
Non-thermal View
One of the first misconceptions of high-pressure is the view as a non-thermal technology. The term non-thermal was coined by Professor
Dietrich Knorr in the 1990s, and in a broad sense it refers to food preservation processes that do not use heat. The view of high-pressure
CO

as a non-thermal technology is clearly at odds with the pressure-volume-temperature relations, Eq. (3). The physical meaning of Eq. (3) is
the heat of compression, which has been investigated quite extensively elsewhere (Patazca et al., 2007; Rasanayagam et al., 2003). The
estimation of the heat of compression can be theoretically calculated according to Eq. (9):

(9)
UN

where α is the thermal conductivity, Cp is the heat capacity, and ρ is the density of the product. Eq. (9) is valid within a small range of
pressures, where the values of α, Cp, and ρ remain unchanged with pressure. Unfortunately, these parameters significantly change within
the domain of high-pressure (400–600 MPa). Moreover, the magnitude of the temperature rise depends on the initial temperature, the
applied pressure, the composition of the food, and the transmitting fluid. Over the years, the estimation of the heat of compression has been
conducted in a number of food products using regression models. Among the most popular models are those based on polynomial equations
of second order. The parameters of such models do not have physical meaning, and they are valid only within the experimental conditions
and the food material from which were derived. A reliable and straightforward approximation to calculate the heat of compression was
developed by Patazca et al. (2007), Eq. (10)
12 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

(10)

where δs is the temperature increase, Tf is the final temperature, and Ti is the initial temperature. All these approaches yield numerical values
of the heat of compression which highlights the flowing of thermal energy. As seen from the brief overview presented here, any claims

F
that high-pressure processing does not use heat are nothing more than a misconception. It is worth mentioning that the temperature (up to
40 °C) used in the industrial pasteurization by high-pressure will not inactivate microorganisms. Thus, pressure is driving force behind the

OO
inactivation rather than a combination of temperature and pressure.

Pressure Versus Temperature Effect


Pressure is a basic thermodynamic variable just like temperature, and they are mathematically related according to Gibbs's definition of
free energy. Gibbs energy is defined by entropy, internal energy, pressure-volume work, and temperature (Eq. 11).

PR
(11)

In a strict sense, during HPP, the effects of temperature cannot be separated from the effects of pressure. This is because, for every
temperature, there is a corresponding pressure. A graphical representation of the pressure and temperature relationship is the water phase
diagram, where the physical state of water is given by a pair of pressure and temperature (Fig. 7). Thermal effects during pressure
treatment can cause volume and energy changes. On the other hand, pressure primarily affects the volume of the product being processed.
The combined net effect during high-pressure processing has been phenomenologically classified as synergistic, antagonistic or additive.

D
Therefore, changes in food such as reactions, phase transitions, molecular reorientation, and others depend on both temperature and pressure
and cannot be treated separately.
TE
Process Uniformity
Ideally, the uniformity of high-pressure requires that the target pressure and temperature are set immediately and applied uniformly
throughout the vessel and product irrespectively of the holding time, product size, and geometry. The view of high-pressure as a
EC
RR
CO
UN

Figure 7 Water phase diagram as influenced by pressure-thermal effects.


High-Pressure Processing: Fundamentals, Misconceptions, and Advances 13

uniform process has been challenged over the past few years (Denys et al., 2000; Grauwet et al., 2016; Nair et al., 2016; Rauh et al.,
2009; Salvi et al., 2017). In this section, the uniformity with respect to pressure and temperature is briefly discussed. The application of
pressure is commonly assumed to be uniform throughout the vessel and product, hydrostatic principle, which is a safe assumption for liquids
and relative homogeneous solids. However, gradients of pressure have been reported for solid foods with hard inclusions (Minerich and
Labuza, 2003; Nair et al., 2016). Minerich and Labuza (2003) reported a pressure gradient of 9 MPa in the center of ham with respect
to the transmitting fluid. Numerical simulations of a model food made of a gel and a wood inclusion showed that hydrostatic pressure
decreased at the interphase between the solid and the inclusion, creating shear stress. In summary, the uniformity of pressure is strongly

F
dependent on the heterogeneity of the food material and the compressibility associated with it. Thus, pressure gradients may exist over a
short time, and the magnitude of such gradients are too small for impacting the overall uniformity of pressure. Contrary, the nonuniformity
of temperature has been documented over the years. The temperature gradients during high-pressure are due to the heat of compression of

OO
the different components involved. Upon compression, the temperature of the transmitting fluid and the food increases. Concomitantly, the
heat of compression for the vessel (steel) is negligible, which creates a thermal gradient. Such a thermal gradient can lead to subsequent
cooling of the sample due to heat loss toward the vessel. The determination of temperature gradients has been performing mainly three
methods (Grauwet et al., 2016):
i) Direct measure of the temperature at different locations.
ii) Computational thermal fluid dynamic modelling (CTFD).

PR
iii) Pressure-temperature-time indicators.

The direct measurement of the thermocouple during compression was the first indicator of the existence of thermal gradients. However,
proper evaluation of the temperature distribution requires multiples thermal sensors across the vessel, which is a daunting task for large
vessels. In addition, wired thermocouples such as Type K and J involve special attention to the sealing of the thermocouple passage through
the vessel (Grauwet et al., 2016). An important limitation of the wired thermocouples is that such devices may affect the normal movement
of the flow inside the vessel, impacting the temperature gradient itself.

D
Numerical simulations of the temperature distribution inside the vessel have been documented elsewhere (Hartmann et al., 2003;
Hartmann et al., 2004; Rauh et al., 2009; Rauh and Delgado, 2010). Rauh et al. (2009) simulated the temperature uniformity in a
vertically oriented vessel under four hypothetical operating scenarios. The first scenario consists of the temperature distribution when
TE
the vessel and the fluid have the same initial temperature, 50 °C. The second hypothetical scenario was considered 20 °C as the initial
temperature for the vessel and the fluid. The third scenario involves the boundary conditions of the first situation but with an increased
viscosity as a way to suppress convection inside the vessel. The last scenario used an initial temperature of 50 °C and a final temperature
equal to the temperature the product after compression under adiabatic conditions (70 °C). For the first scenario, the wall temperature
showed to be colder than the center of the vessel after compression. Thus, heat transfer occurs between the walls and fluid, developing
EC

a temperature and density layer in the vicinity of the walls and grows into the bulk liquid. In the first two scenarios, cold fluid flows
down at the lateral wall, and thermal stratification develops in the vertical direction from bottom to top. At the top surface of the vessel,
unstable temperature stratification arises with colder liquid above warmer liquid. The thermal layering is less developed in scenario three
due to the higher viscosity, where the free convection is suppressed, and only conduction governs the heat transfer. In the last scenario, the
zero-temperature difference between the wall and the bulk fluid leads to a homogeneous temperature field at the end of the pressure-holding
time. Smith et al. (2014) simulated the temperature profiles of a horizontal and vertical vessel and it was demonstrated that the horizontal
RR

arrangement exhibited more temperature inhomogeneity than the vertical vessel.


The development of process indicators has been a topic of research and commercial interest. Indicators such as
pressure-temperature-time can be used as sensors to obtain insight in the temperature gradients within the vessel (Van der Plancken
et al., 2008). A series of protein-based indicators has been developed as a robust system for detecting temperature nonuniformity in
high-pressure vessels (Grauwet et al., 2009, 2010a, 2010b, 2011). These indicators consist of flexible microtubes filled a buffer solution of
a biologically active protein, whose activity after treatment can be easily evaluated spectrophotometrically. Although such indicators can be
CO

used as uniformity test, profiling the temperature distribution within the vessel using these type of indicators presents some limitations. For
instance, indicators are designed for a specific application for which the reaction rate constant is high enough to be detected. Measurements
outside the specific operating conditions significantly impact the reliability of the sensor. Moreover, indicators respond to a specific
temperature range, and uniform readouts at different coordinates inside the vessel do not necessarily prove that no temperature difference
existed (Grauwet et al., 2016).
Overall, temperature gradients within the vessel will exist during compression, and strategies has been developed to minimize the heat
UN

transfer inside the vessel. Selecting components with comparable heat of compression, equilibrating the initial and working temperature,
optimizing the compression rate, using insulated container, and the overall configuration of the hydraulic systems (preventing the injection
of cold transmitting fluid) are examples of such strategies.

Reaction Volume Versus Activation Volume


Over the years, a number of benefits of high-pressure have been documented that deal with changes in the physical properties, effects
on the equilibrium processes, and influences the rate of processes. These processes include crystallization of fat (Zulkurnain et al.,
2019), phase transition of starch (Buckow et al., 2007), ionization equilibria (Stippl et al., 2005), inactivation of enzymes (Hendrickx
et al., 1998), denaturation of proteins, inactivation of microorganisms (Daryaei et al., 2016), mitigate the toxicological potential
of foods (Sevenich et al., 2015), and delaying the rate of some chemical reactions (Martinez-Monteagudo and Saldaña, 2014).
14 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

These applications have been somewhat studied using kinetic analysis with the idea to estimate relevant kinetic parameters (pre-exponential
factor, activation, energy, activation volume, constant rate, and reaction volume). In this section, an attempt to clarify the difference
between reaction volume and activation volume is presented.
Kinetics of chemical reactions as a function of pressure at isothermal conditions is governed by van't Hoff like equations:

(12)

F
OO
(13)

where K and k are the equilibrium constant and the rate constant, respectively; ΔV° is the reaction volume and ΔV≠ is the activation volume.
More specifically, ΔV° is the difference between total partial molar volumes of products and those of reagents. Reaction volume is an
important relationship to determine the effect of a pressure change on the equilibrium constant in dilute solution. ΔV° is somewhat analogous
to the treatment of an ideal gas equilibrium.

PR
Pressure and its corresponding conjugated pair, volume, reveal information on the volume profile of the process. For a hypothetical
reaction (Scheme 1), the reaction volume defined by Eq. (14) and the activation volume according to Eq. (15):

(14)

D
(15)

The measurement of concentrations of species involved in an equilibrium as a function of pressure yields the value of the equilibrium
TE
constant. The change of equilibrium constant with pressure will be positive or negative according to whether the volume change for the
reaction is negative or positive, respectively (Fig. 8). Pressure will, therefore, favor a process that involves a diminution in volume, and
conversely. The determination of the change in volume can be carried out in a dilatometer, where a series of concentrations is obtained.
The subject of measuring reaction volume has been thoroughly reviewed in textbooks and data compilation, where approximate some 1500
reaction volumes of organic and inorganic reaction are collected.
The relative short processing times and the complexity of food matrices make it difficult to reach equilibrium for a given chemical
EC

reaction. Food matrices are very complex systems with various components that simultaneously undergo chemical transformations at
different rates. Consequently, the rate at which the reactions occur is often more important than the equilibrium and therefore most of
RR

Scheme 1
CO
UN

Figure 8 Effect of pressure on the equilibrium constant for a hypothetical reaction.


High-Pressure Processing: Fundamentals, Misconceptions, and Advances 15

the literature considers the activation volume as an indicator whether a particular reaction is promoted or inhibited by pressure. The ΔV≠ is
a reflection of all volume changes that might occur during the progression of a reaction from the ground to the transition state and within
the transition state. Three factors determine the reaction volume of a dissolved species: (i) the intrinsic size of the species as determined
by its van der Waals radius; (ii) the interaction of the species with the solvent to cause electrostriction; (iii) the interaction of the species
with all the solute species. The ΔV° represents those changes in molar volume due to molecular reorganization, alteration of bond angles
and length (known as intrinsic changes, ), and those changes due to interactions of the reactants with the solvent (known as medium
or solvent changes, ), according to Eq. (16).

F
(16)

OO
The term represents the solvent effects, such as changes in polarity, electrostatic forces, dipole interactions, and solvophobic
interactions. If is zero, the reaction is isopolar and the effect of pressure is less pronounced. On the other hand, if is different
from ΔV≠, the medium plays a critical role in the reaction kinetics. The ΔV≠ cannot be measured experimentally because the activated state
is very unstable. Nevertheless, a common practice for estimating the ΔV≠ is to evaluate the effect of pressure on the constant reaction rate
at isothermal conditions in the form of Eq. (17).

PR
(17)

Most important is that Eq. (17) explains the effect of pressure on the overall activation volume, which accounts for the changes in
volume occurring during the transition state and interaction of reactants with the solvent. This is somewhat analogous to the apparent
activation energy reported for some food systems. Examples of activation volumes on food systems are reviewed elsewhere

D
(Martinez-Monteagudo and Saldaña, 2014). TE
Applications

The motivation of using high pressure lies on chemical and physical effects induced by pressure. There are three important consequences
of applying high hydrostatic pressure:
1) changes in physical properties, such as melting point, solubility, density, viscosity, etc.
EC

2) effects on equilibrium processes, such as acid-base equilibria, ionization, etc.


3) effects on rates of processes, such as delaying or accelerating the rate at which a particular reaction occurs.

In general, qualities attributes such as microbiological safety, functionality, instrumental quality and nutritional attributes are the
resultants of the way of these three phenomena are affected by pressure. For instance, inactivation of microorganisms is a combination of
changes in physical properties of membrane lipids, changes in the chemical equilibrium that modify the internal pH, and changes in the
RR

rate of specific physiological functions that cause irreversible or lethal damage on bacteria cells (Molina-Gutierrez et al., 2002). Table 5
summarizes various examples on how food quality attributes are dictated by the way of pressure affect the physical, equilibrium and rate
processes. A rate process was considered when pressure increase or decrease the concentration of a particular compound.

Pasteurization
CO

Fig. 9 exemplifies a hypothetical flow diagram of a high-pressure operation. The process starts with traditional operations such as blending,
mixing, or dissolving, depending on the product to be treated. Then, the food to be treated is preferably vacuum packaged in a flexible
polymeric package, and subsequently loaded inside a cylindrical carrier basket. Then, the carrier basket containing the packed product is
loaded into the stainless-steel vessel. Subsequently, the pressure vessel is closed with the end closures. The pressure vessel is filled with
pressure transmitting fluid. The target process pressure is achieved through compression of pressure transmitting fluid using the combined
action of a pump and intensifier. Once the desired pressure is obtained, the product is held for the desired time. After this time, the vessel
UN

depressurized quickly and the product is unloaded.


High hydrostatic pressure has been effective in inactivating variety of pathogenic and spoilage vegetative cells, yeast, mold, and
viruses. The inactivation level depends on the pressure used (typically between 400 and 600 MPa), temperature (25–50 °C), time (3–10
min), the composition of the food, pH, and water activity. The microbial effectiveness of HPP depends on the type of microorganisms.
In other words, the sensitivity towards HPP may vary from different microorganisms and strains (Balci and Wilbey, 1999). Mackey
et al. (1995) reported that high-pressure caused physical changes to the cell structure, stationary phase cells are more resistant than
the exponential phase cells. These authors also reported that treatment of 400 MPa for 10 min caused a log reduction of 1.3 in
Listeria monocytogenes cell at the stationary phase, while the exponential phase was reduced by more than 7.0 log. The resistance of
certain microorganisms towards pressure is related to the morphology of the cells. Rod-shaped bacteria are most sensitive, while the
spherical shaped are considered one of the most resistant. Yeasts and molds that cause spoilage in foods are very sensitive to HPP.
16 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

Table 5 Examples on how food quality attributes are influenced by the effect of pressure on physical properties, equilibrium and rate processes

Quality attribute Effect upon pressure treatment

Physical property Equilibrium process Rate process

Mechanically shucking of - Detaching the - Releasing intramuscular - The extend of


oysters adductor muscle components shucking depends on

F
the pressure level
Improving of meat - Pressure causes - Shifts the pH value that - Biochemical

OO
tenderness by pressure- disaggregation prevents re-association of reactions are
heat treatment of proteins proteins fragments controlled by
pressure
Starch gelatinization - Changes in - Shifts the chemical balance - Retrogradation rate
viscosity of is controlled by
starch pressure
suspension

PR
- Changes in
rheological
properties
Increasing cheese yield - Disruption of - Alter mineral balance - Pressure controls the
casein micelles (colloidal and soluble extend of protein-fat
- Denaturation of calcium) interactions
whey proteins
Improving texture in egg - Increase - Change in pH, which - Pressure affects the

D
patties viscosity of egg enhances hydrophobic rate of protein
patties interactions aggregation
- Protein
TE
aggregation
Cell membrane damage in - Induce - Shifts the balance between - Pressure controls
onions membrane extra- and intra-cellular pH diffusion of solutes
permeabilization within the cell

Reprinted by permission from Springer Nature: Food Engineering Reviews Martinez-Monteagudo and Saldaña (2014).
EC
RR
CO
UN

Figure 9 Schematic diagram of a hypothetical high-pressure processing operation.


High-Pressure Processing: Fundamentals, Misconceptions, and Advances 17

Several researchers have demonstrated the effectiveness of HPP for pasteurizing liquid and solid foods such as deli meats, salads, seafood,
fruit juices, and vegetable products (Balasubramaniam et al., 2015).

Pressure-Assisted Thermal Processing


The application of high-pressure can be carried out without or with the application of external heating. In the first case, the operation
is known as high-pressure processing (HPP), and it involves the use of pressure levels in the range of 100–800 MPa at near room

F
temperature in order to inactivate pathogenic and spoilage vegetative bacteria, yeast, and molds. HPP results in a pasteurization effect.
Since pasteurization does not inactivate bacterial spores, it is important to maintain the pressure pasteurized product under refrigerated
storage and handling.

OO
On the other hand, the combination of pressure and heating is known as Pressure-assisted thermal processing (PATP). This technology
involves the preheating of food materials to about 75–90 °C, and subsequently the application of 600 MPa for 5–10 min. The rapid
temperature increase during hydrostatic compression and subsequent cooling upon decompression is a unique feature of PATP.
Interestingly, PATP is somewhat analogous to in-container sterilization (retorts), where the food is packaged, loaded, and treated for a
prescribed combination of time-temperature or time-pressure. Despite such similarities, their governing principles are quite different. Fig.
10 attempts to provide a detailed description of a typical PATP treatment. The treatment time is the sum of the loading (t1), compression (t2),

PR
holding (t3), decompression (t4), and unloading (t5) time. At the beginning of the treatment, a drop in the medium temperature (indicated
by arrow (1)). This drop occurs when the preheated sample is inserted in the high-pressure vessel. Arrow (1) also indicates the start of
loading time, which is the time needed to insert the preheated sample, adjust the transmitted fluid volume and close the high-pressure
vessel. Then, the sample is pressurized until it reaches the target pressure. This period is known as compression or build-up time. Due to
the adiabatic heating, the temperature of both sample and medium rises (arrow (2)). During t2, the sample is submitted to a non-isothermal
and non-isobaric conditions over a relatively short time. At the point at which the target temperature and pressure have been reached
is considered as the start of the holding time. The holding time is the only period at which the sample is at isothermal and isobaric

D
conditions. At the end of the holding time, the decompression takes place which is characterized by a drop in the medium temperature
indicated by arrow (3). This period is usually short and indicates the end of the treatment. In 2009, FDA approved an industrial petition
for sterilization of mashed potato by PATP. More recently, Pressure Enhanced Sterilization (PES) received regulatory acceptance for
TE
commercial production of multi-component foods. Although there are no low-acid products preserved by PATP commercialized yet, this
technology has the potential to deliver a variety of low-acid products, such as egg, and milk-based products, baby foods, vegetables,
ready-to-eat foods, desserts, gravies, soups, and sauces. Further, units for PATP are primarily restricted to pilot scale (5–55 L). Since PATP
utilize intensive pressure and heat, from the standpoint of material science and engineering, the process demands significant stress on the
vessel and seals, potentially limiting the equipment life. Additional PATP technology limitation is the preheating process, wherein food
EC

packages are typically heated (75–90 °C) by conventional conduction and convection heat transfer at ambient pressures.
RR
CO
UN

Figure 10 Illustration of a typical pressure-assisted thermal processing treatment. t1−5 is the floading, compression, holding, decompression, and
unloading time. Figure adapted from Martínez-Monteagudo et al. (2012), with permission from Elsevier.
18 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

Operational Mode
High-pressure processing of foods is fundamentally a batch operation, wherein a prepacked product is subjected to a given pressure
(400–600 MPa) for a specific time (<5 min). Although the actual application of pressure may take a few minutes, the entire operation
requires much longer time. The time for typical HPP consists the time needed for filling the vessel with the transmitting fluid, loading
the sample, adjusting the final volume, closing the vessel, pressurization, holding, depressurization, opening, and unloading the sample.
Another important feature of HPP regarding its operational mode is the size of the vessel which is limited not only by the strength of
the construction material but also by the availability of the large parts of high-grade steel. Increasing the throughput of HPP involves

F
the repetition of all operational steps carried out during the batch process. Thus, the development and implementation of continuous
high-pressure processing would contribute to the cost-effectiveness and scalability of the technology. Fig. 11 shows an arrangement for

OO
continuous high-pressure processing. The working principle consists of using a pumpable food as a transmitting fluid in the high-pressure
chamber of the intensifier, while the fluid acting in the low-pressure chamber of the intensifier is a typical transmitting fluid. The product
to be treated is delivered by a low-pressure pump into the vessel, which is close by a set of check valves. The target pressure inside the
vessel is accomplished by means of a high-pressure pump and an intensifier. After the treatment, the liquid is transferred to a balance tank
which can be connected to a filling machine. This arrangement offers the advantage to increase the throughput by connecting more vessels
in such a way that while one vessel is being filled, another one is pressurized and a third one is being emptied. Although conceptual process
design has been developed for semi-continuous process, the industrial implementation of HPP remains a batch operation (Lelieveld and

PR
Hoogland, 2016).

High-Pressure Freezing and Thawing


Pressure levels of 200 MPa cause a reversible depression of the freezing point of water 0 to −21 °C. Such a phenomenon has been exploited
to rapidly freeze and thaw g high moisture content foods, known as pressure-assisted freezing (PAF) and pressure-assisted thawing (PAT),

D
respectively. During PAF, the samples are cooled under pressure up to its phase change temperature at applied pressure. The product
is frozen under pressure by supercooling at faster ice-nucleation rate. This process helps in preserving the microstructure of food and
biological materials. Pressure-assisted thawing (PAT) involves thawing food material under constant pressure. The process can help in
reducing the thawing time and drip loss. Another potential application is the storage at freezing conditions (up to −20 °C) without the
TE
transition from liquid to solid (You et al., 2016).
EC
RR
CO
UN

Figure 11 Schematic representation of a semi-continuous high-pressure processing. Reprinted by permission from Elservier: Innovative Food
Science & Emerging Technologies Martinez-Monteagudo et al. (2012).
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 19

Pressure-Assisted Extraction
Extraction of valuable compounds from biological matrix has been enhanced by high hydrostatic pressure. Extraction yields obtained by
pressure mainly depend on the pressure, time, and type solvent used. Mixtures of organic solvents have been tested under pressure to
assist the extraction of antioxidants and other bioactive compounds. Upon decompression, pressure can break weak chemical bonds of the
extracting matrix, making some compounds available for extraction.

F
Conclusions

OO
As seen in this chapter, high-pressure has been evolving very successfully. Over the past few years, multiple applications have appeared,
while the old ones have expanded. Perhaps the most important development is the commercialization of pasteurized products by
high-pressure. This is because high-pressure provides a unique opportunity to food processors for inactivating various harmful pathogenic
and spoilage microbes without adversely impacting product quality and nutritional attributes. A pseudo-instantaneous temperature increase
during compression, and subsequent cooling upon decompression, substantially reduce the thermal damage. Extended shelf-life products
and pressure-sterilized products are likely to follow this trend. High-pressure units can be seen as unit operation, where theory and practice
can combine to yield a desirable modification within the food processing line. It is hoped that the discussion of some misconceptions will

PR
aid in a better understanding of the development of high-pressure.

Acknowledgments

This work has been made possible through the partial financial support of USDA National Institute for Food and Agriculture (HATCH

D
project SD00H607-16).

References
TE
Balasubramaniam, V.M., Barbosa-Canovas, G.V., Lelieveld, H.L.M., 2016. High-pressure processing equipment for the food industry. In: Balasubramaniam,
V.M., Barbosa-Canovas, G.V., Lelieveld, H.L.M. (Eds.), High Pressure Processing of Food Principles, Technology and Applications. Springer
Science+Business, pp. 39–66.
Balasubramaniam, V.M., Martínez-Monteagudo, S.I., Gupta, R., 2015. Principles and application of high pressure–based Technologies in the food industry.
Ann. Rev. Food Sci. Technol. 6 (1), 435–462.
EC

Balci, A.T., Wilbey, R.A., 1999. High pressure processing of milk‐the first 100 years in the development of a new technology. Int. J. Dairy Technol. 52 (4),
149–155.
Bridgman, P.W., 1909. An experimental determination of certain compressibilities. Proc. Am. Acad. Arts Sci. 44 (10), 255–279.
Bridgman, P.W., 1914a. Change of phase under pressure. I. The phase diagram of eleven substances with especial reference to the melting curve. Phys. Rev.
3 (2), 126–141.
Bridgman, P.W., 1914b. The coagulation of albumen by pressure. J. Biol. Chem. 19 (4), 511–512.
RR

Bridgman, P.W., 1923. The thermal conductivity of liquids under pressure. Proc. Am. Acad. Arts Sci. 59 (7), 141–169.
Bridgman, P.W., 1935. Theoretically interesting aspects of high pressure phenomena. Rev. Mod. Phys. 7 (1), 1–33.
Brown, K., 1920. The manufacture of phenol in a continuous high pressure autoclave. J. Ind. Eng. Chem. 12 (3), 279–280.
Buckow, R., Heinz, V., Knorr, D., 2007. High pressure phase transition kinetics of maize starch. J. Food Eng. 81 (2), 469–475.
Butz, P.B., Serfert, Y., Garcia, A.F., Dieterich, S., Lindauer, R., Bognar, A., Tauscher, B., 2004. Influence of high-pressure treatment at 25°C and 80°C on
folates in orange juice and model media. J. Food Sci. 69 (3), SNQ117–SNQ121.
Daryaei, H., Yousef, A.E., Balasubramaniam, V.M., 2016. Microbiological aspects of high-pressure processing of food: inactivation of microbial vegetative
cells and spores. In: Balasubramaniam, V.M., Barbosa-Cánovas, G.V., Lelieveld, H.L.M. (Eds.), High Pressure Processing of Food: Principles,
CO

Technology and Applications. Springer New York, New York, NY, pp. 271–294.
Denys, S., Van Loey, A.M., Hendrickx, M.E., 2000. A modeling approach for evaluating process uniformity during batch high hydrostatic pressure
processing: combination of a numerical heat transfer model and enzyme inactivation kinetics. Innov. Food Sci. Emerg. Technol. 1 (1), 5–19.
Dewar, J., 1882. The critical point of mixed vapours. Proc. R. Soc. 30, 538–546.
Eggers, R., 2005. Historical retrospect on high-pressure processes. In: Eggers, R. (Ed.), Industrial High Pressure Applications: Processes, Equipment and
Safety. Wiley-VCH Verlag GmbH & Co. KGaA, pp. 1–46.
Eyring, H., 1935. The activated complex and the absolute rate of chemical reactions. Chem. Rev. 17 (1), 65–77.
UN

Grauwet, T., Plancken, I.V.d., Vervoort, L., Hendrickx, M.E., Loey, A.V., 2010a. Mapping temperature uniformity in industrial scale HP equipment using
enzymatic pressure–temperature–time indicators. J. Food Eng. 98 (1), 93–102.
Grauwet, T., Van der Plancken, I., Verbeyst, L., Hendrickx, M., van loey, A., 2016. High-pressure processing uniformity. In: Balasubramaniam, V.M.,
Barbosa-Canovas, G.V., Lelieveld, H.L.M. (Eds.), High Pressure Processing of Food: Principles, Technology and Applications. Springer
Science+Business Media, New York.
Grauwet, T., Van der Plancken, I., Vervoort, L., Hendrickx, M.E., Loey, A.V., 2009. Investigating the potential of Bacillus subtilis α-amylase as a
pressure-temperature-time indicator for high hydrostatic pressure pasteurization processes. Biotechnol. Prog. 25 (4), 1184–1193.
Grauwet, T., Van der Plancken, I., Vervoort, L., Hendrickx, M.E., Van Loey, A., 2010b. Solvent engineering as a tool in enzymatic indicator development for
mild high pressure pasteurization processing. J. Food Eng. 97 (3), 301–310.
Grauwet, T., Van der Plancken, I., Vervoort, L., Matser, A., Hendrickx, M., Van Loey, A., 2011. Temperature uniformity mapping in a high pressure high
temperature reactor using a temperature sensitive indicator. J. Food Eng. 105 (1), 36–47.
Hartmann, C., Delgado, A., Szymczyk, J., 2003. Convective and diffusive transport effects in a high pressure induced inactivation process of packed food. J.
Food Eng. 59 (1), 33–44.
20 High-Pressure Processing: Fundamentals, Misconceptions, and Advances

Hartmann, C., Schuhholz, J.-P., Kitsubun, P., Chapleau, N., Le Bail, A., Delgado, A., 2004. Experimental and numerical analysis of the thermofluiddynamics
in a high-pressure autoclave. Innov. Food Sci. Emerg. Technol. 5 (4), 399–411.
Hendrickx, M., Ludikhuyze, L., Van den Broeck, I., Weemaes, C., 1998. Effects of high pressure on enzymes related to food quality. Trends Food Sci.
Technol. 9 (5), 197–203.
Hite, B.H., 1899. The effect of high pressure preservation of milk. W. Va. Agric. Exp. Station Bull. 58, 15–35.
Indrawati, Arroqui, C., Messagie, I., Nguyen, M.T., Van Loey, A., Hendrickx, M., 2004. Comparative study on pressure and temperature stability of
5-methyltetrahydrofolic acid in model systems and in food products. J. Agric. Food Chem. 52 (3), 485–492.
Illinois Institute of Technology's Institute for Food Safety and Health, 2015. IFSH receives FDA acceptance of pressure enhanced sterilization process for
commercial production of MultiComponent shelf-stable foods. Food Safety Mag.

F
Kalichevsky, M.T., Knorr, D., Lillford, P.J., 1995. Potential food applications of high-pressure effects on ice-water transitions. Trends Food Sci. Technol. 6,
253–259.
Kennick, W.H., Elgasim, E.A., Holmes, Z.A., Meyer, P.F., 1980. The effect of pressurisation of pre-rigor muscle on post-rigor meat characteristics. Meat Sci.

OO
4 (1), 33–40.
King, J.W., 2014. Modern supercritical fluid technology for food applications. Ann. Rev. Food Sci. Technol. 5 (1), 215–238.
Larson, W.P., Hartzell, T.B., Diehl, H.S., 1918. The effect of high pressures on bacteria. J. Infect. Dis. 22 (3), 271–279.
Lelieveld, H.L.M., Hoogland, H., 2016. Continuous high-pressure processing to extend product shelf life. In: Balasubramaniam, V.M., Barbosa-Canovas,
G.V., Lelieveld, H.L.M. (Eds.), High Pressure Processing of Food. Springer Science+Business.
Mackey, B., Forestiere, K., Isaacs, N.S., 1995. Factors affecting the resistance of Listeria monocytogenes to high hydrostatic pressure. Food Biotechnol. 9
(1–2), 1–11.
Martínez-Monteagudo, S.I., Balasubramaniam, V.M., 2016. Fundamentals and applications of high-pressure processing technology. In: Balasubramaniam,

PR
V.M., Barbosa-Canovas, G.V., Lelieveld, H.L.M. (Eds.), High Pressure Processing of Food. Springer Science+Business, New York, NY, pp. 3–17.
Martinez-Monteagudo, S.I., Saldaña, M.D.A., 2014. Chemical reactions in food systems at high hydrostatic pressure. Food Eng. Rev. 6 (4), 105–127.
Martínez-Monteagudo, S.I., Saldaña, M.D.A., Torres, J.A., Kennelly, J.J., 2012. Effect of pressure-assisted thermal sterilization on conjugated linoleic acid
(CLA) content in CLA-enriched milk. Innov. Food Sci. Emerg. Technol. 16, 291–297.
Mathys, A., Knorr, D., 2009. The properties of water in the pressure–temperature landscape. Food Biophys. 4, 77–82.
Matser, A.M., Krebbers, B., van den Berg, R.W., Bartels, P.V., 2004. Advantages of high pressure sterilisation on quality of food products. Trends Food Sci.
Technol. 15 (2), 79–85.
Minerich, P.L., Labuza, T.P., 2003. Development of a pressure indicator for high hydrostatic pressure processing of foods. Innov. Food Sci. Emerg. Technol.

D
4 (3), 235–243.
Molina-Gutierrez, A., Stippl, V., Delgado, A., Gänzle, M.G., Vogel, R.F., 2002. In situ determination of the intracellular pH of Lactococcus lactis and
Lactobacillus plantarum during pressure treatment. Appl. Environ. Microbiol. 68 (9), 4399–4406.
Mozhaev, V.V., Heremans, K., Frank, J., Masson, P., Balny, C., 1994. Exploiting the effects of high hydrostatic pressure in biotechnological applications.
TE
Trends Biotechnol. 12 (12), 493–501.
Murzin, D., Salmi, T., 2005. Elementary reactions. In: Catalytic Kinetics. Elsevier Science, Amsterdam, (Chapter 3).
Nair, A., Maldonaldo, J.A., Miyazawa, Y., Cuitiño, A.M., Schaffner, D.W., Karwe, M., 2016. Numerical simulation of stress distribution in heterogeneous
solids during high pressure processing. Food Res. Int. 84, 76–85.
Pandolfe, W.D., 1982. Development of the new Gaulin micro-gap™ homogenizing valve. J. Dairy Sci. 65 (10), 2035–2044.
Park, S.H., Balasubramaniam, V.M., Sastry, S.K., 2014. Quality of shelf-stable low-acid vegetables processed using pressure–ohmic–thermal sterilization.
EC

LWT – Food Sci. Technol. 57, 243–252.


Patazca, E., Koutchma, T., Balasubramaniam, V.M., 2007. Quasi-adiabatic temperature increase during high pressure processing of selected foods. J. Food
Eng. 80 (1), 199–205.
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. J. Food Sci. 68 (1), 254–259.
Rauh, C., Baars, A., Delgado, A., 2009. Uniformity of enzyme inactivation in a short-time high-pressure process. J. Food Eng. 91 (1), 154–163.
Rauh, C., Delgado, A., 2010. Analytical considerations and dimensionless analysis for a description of particle interactions in high pressure processes. High
RR

Press. Res. 30 (4), 567–573.


Salvi, D., Khurana, M., Karwe, M.V., 2017. Prediction of temperature distribution in a horizontal high pressure food processing vessel and its impact on
process uniformity. J. Food Process. Eng. 40 (5), e12547.
Sander, F.V., 1943. The effects of high pressure on the inversion of sucrose and the mutarotation of glucose. J. Biol. Chem. 148 (2), 311–319.
Sevenich, R., Mathys, A., 2018. Continuous versus discontinuous ultra-high-pressure systems for food sterilization with focus on ultra-high-pressure
homogenization and high-pressure thermal sterilization: a review. Compr. Rev. Food Sci. Food Saf. 17, 646–662.
Sevenich, R., Bark, F., Kleinstueck, E., Crews, C., Pye, C., Hradecky, J., Reineke, K., Lavilla, M., Martinez-de-Maranon, I., Briand, J.C., Knorr, D., 2015.
CO

The impact of high pressure thermal sterilization on the microbiological stability and formation of food processing contaminants in selected fish systems
and baby food puree at pilot scale. Food Control 50, 539–547.
Sizer, C.E., Balasubramaniam, V.M., Ting, E., 2002. Validating high-pressure processes for low-acid foods. Food Technol.-Chicago 2, 36–42.
Skelton, E.F., Webb, A.W., 2003. High-pressure research. In: Meyers, R.A. (Ed.), Encyclopedia of Physical Science and Technology, third ed. Academic
Press, New York, pp. 345–363.
Smith, N.A.S., Knoerzer, K., Ramos, A.M., 2014. Evaluation of the differences of process variables in vertical and horizontal configurations of High Pressure
Thermal (HPT) processing systems through numerical modelling. Innov. Food Sci. Emerg. Technol. 22, 51–62.
Soxhlet, F., 1881. Supposed conversion of starch into sugar by water at a high temperature. Biedermann's Zentralblatt 86, 554–557.
UN

Stern, O., 1897. The influence of a pressure of 500 atmospheres on the rate of inversion of cane sugar. Ann. Phys. 86, 554–557.
Stippl, V.M., Delgado, A., Becker, T.M., 2005. Ionization equilibria at high pressure. Eur. Food Res. Technol. 221 (1), 151–156.
Van der Plancken, I., Grauwet, T., Oey, I., Van Loey, A., Hendrickx, M., 2008. Impact evaluation of high pressure treatment on foods: considerations on the
development of pressure–temperature–time integrators (pTTIs). Trends Food Sci. Technol. 19 (6), 337–348.
Verbeyst, L., Oey, I., Van der Plancken, I., Hendrickx, M., Van Loey, A., 2010. Kinetic study on the thermal and pressure degradation of anthocyanins in
strawberries. Food Chem. 123 (2), 269–274.
Vetter, G., 2001. In: In: Bertucco, A., Vetter, G. (Eds.), High Pressure Process Technology: Fundamentals and Applications, vol. 9, Elsevier Science.
Wanekaya, A.K., Myung, S., Sadik, O.A., 2002. Pressure assisted chelating extraction: a novel technique for digesting metals in solid matrices. Analyst 127,
1272–1276.
Yan, B., Martínez-Monteagudo, S.I., Cooperstone, J.L., Riedl, K.M., Schwartz, S.J., Balasubramaniam, V.M., 2017. Impact of thermal and pressure-based
Technologies on carotenoid retention and quality attributes in tomato juice. Food Bioprocess Technol. 10 (5), 808–818.
You, J., Habibi, M., Rattan, N., Ramaswamy, H.S., 2016. Pressure shift freezing and thawing. In: Balasubramaniam, V.M., Barbosa-Cánovas, G.V., Lelieveld,
H.L.M. (Eds.), High Pressure Processing of Food: Principles, Technology and Applications. Springer New York, New York, NY, pp. 143–166.
High-Pressure Processing: Fundamentals, Misconceptions, and Advances 21

Zulkurnain, M., Balasubramaniam, V.M., Maleky, F., 2019. Effects of lipid solid mass fraction and non-lipid solids on crystallization behaviors of model fats
under high pressure. Molecules 24 (15), 2853.

F
OO
PR
D
TE
EC
RR
CO
UN

View publication stats

You might also like