You are on page 1of 10

Journal of Food Engineering 166 (2015) 45–54

Contents lists available at ScienceDirect

Journal of Food Engineering


journal homepage: www.elsevier.com/locate/jfoodeng

Rheological properties and microstructure of tomato puree subject


to continuous high pressure homogenization
J. Tan, W.L. Kerr ⇑
Department of Food Science and Technology, University of Georgia, Athens, GA 30602, United States

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

Article history: Tomato puree was processed by continuous high-pressure (CHP) homogenization at 69–276 MPa, for 1–3
Received 28 January 2015 passes. Laser scattering and light microscopy showed CHP reduced the pulp particles to 10–100 lm,
Received in revised form 6 April 2015 producing smaller and more uniform particles, with processing at 276 MPa and 2 passes producing great-
Accepted 18 May 2015
est particle reduction. No differences in moisture or color were found due to different treatments. In gen-
Available online 19 May 2015
eral, G0 > G00 for all samples, suggesting a soft gel network. Both the storage modulus (G0 ) and loss modulus
(G00 ) decreased with CHP pressure. G0 decreased modestly with frequency between 0.1 and 2 Hz, and more
Keywords:
dramatically between 2 and 30 Hz, with behavior characteristic of entangled polymers. In general, yield
Tomato puree
High pressure
stress decreased with homogenization pressure, but increased with number of passes. CHP-treated sam-
Rheological properties ples had lower consistency and were less shear-thinning than the control. Repeated passes increased the
consistency of CHP samples. The results suggest CHP processing produced smaller and more uniform par-
ticles, causing a reduced level of microstructure that contributes to elastic properties at small
deformation.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction determines the rheological properties of the fluid. In addition,


tomatoes are usually thermally processed during chopping in
Tomatoes are one of the world’s most significant crops and are hot- or cold- break processes. This inhibits enzymes that would
cultivated globally. In 2012, total tomato production was esti- cause the tomato solids to gel or aggregate, thus preventing lumpi-
mated at 161,800,000 metric tons (FAO, 2012). Tomatoes are often ness and promoting homogeneity.
consumed fresh, but several varieties are processed into tomato Products intended for ketchup, purees or sauces may also
sauce, soup, paste, puree, juice, ketchup and salsa. Tomatoes are undergo a homogenization step to further reduce particle size, bet-
a nutritious food with substantial amounts of vitamin C and biotin, ter mix components and enhance product stability. In this context,
as well as other vitamins and minerals. A 100 g serving of tomatoes high-pressure homogenization involves metering the product
typically has 1.5 g of fiber, with 90% of this as insoluble fiber. through a small orifice between the homogenizer valve and valve
Tomatoes also contain several phytonutrients including lycopene seat, at pressures between 10 and 70 MPa. Thakur et al. (1995)
and other carotenoids, as well as flavonols and hydrocinnamic found that the consistency of homogenized tomato juice increased
acids. Several studies have investigated the benefits of tomatoes with pressures up to 3000 PSI (20.7 MPa), although the amount of
in reducing the risk of heart disease, improving bone health, and the change depended on whether the juice was hot- or cold- pro-
decreasing the risk of cancer. The role of tomato products in reduc- cessed. The combined effects of turbulence, cavitation and shear
ing the risk of prostate cancer has received particular attention, help disrupt tissue pieces, forming smaller clusters of cells. This
especially as it relates to lycopene, b-carotene and b-tomatine occurs primarily by separating the pieces along the cell walls
(Giovannucci, 1999; Etminan et al., 2004; Kirsh et al., 2006). (Stang et al., 2001).
To make flowable products, tomatoes are subject to crushing or In recent years, there have been several studies on the effects of
other particle size reduction steps. The crushed tomato is typically very high pressure on the particle size, physical properties and
sent through a pulper and finisher, which are rotary sieves that microbiological status of fruit and vegetable based liquids. In
remove overly large pieces. The size and shape of the resulting hydrostatic high pressure processing (HPP), products are sealed
particles, along with the final solids content of the mixture, in flexible pouches, placed within a sealed chamber, and subject
to pressures of 200–800 MPa as delivered by surrounding water.
For liquid foods, continuous high pressure (CHP) processing has
⇑ Corresponding author. also been studied. This is an extension of conventional

http://dx.doi.org/10.1016/j.jfoodeng.2015.05.025
0260-8774/Ó 2015 Elsevier Ltd. All rights reserved.
46 J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54

homogenization, and uses specialized pumps and valves to achieve tests include small strain ramps to identify yield stresses at which
operating pressures up to 400 MPa. Thus, fluid foods can be pro- the material begins to flow, or dynamic oscillatory tests to study
cessed in a continuous manner. the viscoelastic properties of the colloidal suspension and how
In early studies, hydrostatic pressure was used to process these vary with temperature or frequency (Tabilo-Munizaga and
non-heated tomato juice contained in polyethylene pouches at Barbosa-Cánovas, 2005).
pressures of 500–900 MPa (Porretta et al., 1995). Processed juice A few studies have examined the effects of very high-pressure
had lower counts of bacteria, yeast and molds than control. Juice homogenization on the flow properties of food materials, although
viscosity increased with pressure, and the authors noted that very these have been mostly for formulated emulsions. For example,
high pressures gave rise to a jelly-like translucent structure. While Floury et al. (2002) showed that CHP processing could convert soy-
color was improved, the HPP products were found to be rancid and bean protein-stabilized emulsions from Newtonian liquid emul-
inedible. Hernandez and Cano (1998) found that combined HPP sions into shear-thinning emulsion gels. Lopez-Sanchez et al.
(50–500 MPa) and heating could reduce 32.5% of the pectin methy- (2011a) studied emulsions produced with 5% olive oil and com-
lesterase (PME) in tomato puree, and up to 25% of the peroxidase. minuted carrot/tomato blends. They found that pressures of 10
Overly high pressures, such as 335–500 MPa for PME, caused acti- and 100 MPa, as well as multiple cycles of homogenization, had
vation of enzymes. Others have studied the effects of HPP on the significant effects on both particle size and viscoelastic properties
extractability of carotenoids of tomato pulp puree (Garcia et al., of the blends. Mert (2012) studied high pressure microfluidization
2001). While no effects were found on the extractability of carote- up to 200 MPa and its effect on lycopene in ketchup. That work
noids, water binding was enhanced after HPP and the rate of glu- used lower pressures than presented here, and with a single pass
cose diffusion through dialysis tubing was reduced. Rodrigo et al. through the system. In addition, the microfluidizer uses a substan-
(2007) studied the changes in color of HPP tomato puree, pro- tially different valve system than in this study.
cessed for 60 min at 65 °C and between 300 and 700 MPa. No There has been considerable interest in using physical processes
degradation in color was measured for the combined thermal to modify the properties of food ingredients. CHP homogenization
and high pressure processing. Verlent et al. (2006) measured some has shown promise for altering the size of suspended solid or liq-
rheological properties of tomato homogenates. When processed at uid particles, modifying product rheological properties, and affect-
60 °C, they found that the consistency decreased when processed ing the sensory properties of the food material. The purpose of this
at pressure up to 300 MPa. At higher pressure (500 MPa), the study was to investigate the use of CHP for modifying the structure
homogenates had higher consistency, but did begin to develop a and rheological properties of tomato puree. This included the use
soft-gel translucent structure. Researchers have found that static of single pass homogenization at pressures not previously studied
high pressure treatments did not greatly affect the color of tomato (up to 275 MPa), as well as several passes through the CHP system.
puree or strawberry juice, implying that the lycopene or antho- The rheological properties studied included shear rate dependency
cyanins were not greatly lost during the process (Rodrigo et al., of viscosity, static yield stress and dynamic measurements of stor-
2007). age and loss moduli as a function of frequency. In addition, color
As with traditional homogenization, continuous high-pressure and particle size was assessed, and changes in the tomato tissue
processing contributes substantial shear and turbulence to the during processing were examined through light microscopy.
fluid, and thus provides greater opportunity for tissue disruption.
Corbo et al. (2010) used high-pressure homogenization as a means
to control the growth of molds in tomato juice. They found that the 2. Materials and methods
levels of Fusarium oxysporum decreased somewhat with pressure
between 30 and 150 MPa, and could be eliminated after three 2.1. Materials
passes in the homogenizer. Colle et al. (2010) examined the effects
of high-pressure homogenization (8.4–132 MPa) on tomato pulp Roma tomatoes (Solanum lycopersicum L.) were purchased in
microstructure and the amount and bioaccessibility of lycopene. 23 kg boxes from a local market in Athens, Georgia. The tomatoes
They found that increasing pressure caused a breakdown of tomato were at Stage 6 ripeness (>90% red surface) and stored refrigerated
cell aggregate structures and had no effect on total lycopene, but at 2 °C for less than 2 days prior to processing. Before further pro-
enhanced the in vitro bioaccessibility of lycopene. CHP homoge- cessing, skins were removed by first immersing 11.3 kg batches
nization has also been investigated as a means to form very fine into boiling water contained in a 150 L steam jacketed kettle
emulsions. Qian and McClements (2011) made emulsions of corn (Model M-195, Legion Utensil Inc., Long Island City, NY). The toma-
oil/octadecane in glycerol solutions, using a microfluidizer. They toes were contained within a 35 cm  35 cm  35 cm mesh con-
found that both higher pressure and multiple passes contributed tainer for 8 min and removed once the skins became loose.
to the reduction of particle size, with particles attained on the
order of 116–240 nm. Yuan et al. (2008) were able to obtain parti-
cles of b-carotene in the range 113–168 nm at pressures between 2.2. Preprocessing
60 and 140 MPa. Lopez-Sanchez et al. (2011a) studied the proper-
ties of olive oil emulsions prepared with comminuted carrot and Softened tomatoes were loaded into a paddle pulper-finisher
tomato material. (Model 185S, Langsenkamp Inc., Indianapolis, IN) to remove the
The rheological properties of tomato purees are critical to the seeds and skin. The best recovery was attained with a mesh size
product as they determine sensory properties such as mouthfeel of 1/16 inch (1.58 mm). The discharged pulp was then subject to
and perceived consistency, how easily consumers can cause the a hot-break process to inactivate pectin methylesterase (PME)
puree to flow, and the design of equipment used to process the and polygalacturonase (PG). The pulp was pumped by a rotary
puree (Herch et al., 2000). The rheological properties are deter- pump (Model X5SS1PTYDGHLW, Roper Pump INC, Commerce,
mined by the level of tomato solids (both insoluble and soluble), GA) through a stainless coil heat exchanger (305 cm length,
the particle size and shape of tomato tissue pieces, and the degree 8 mm ID and 10 mm OD) heated by boiling water in a steam kettle.
to which molecules such as pectin or hemicellulose are dissociated The puree reached a temperature of 94 °C, as measured by a K-type
from the cells. In the simplest case, the consistency and flow of thermocouple placed where the product exited the heated coil. The
tomato puree under specified conditions of shear rate, shear stress puree was immediately cooled in a second stainless coil heat
and temperature are of interest (McKenna, 2003). More in depth exchanger immersed in an ice-water bath, allowing it to reach a
J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54 47

temperature of 40 °C. The pump operated at 720 rpm, causing the viscosity versus shear stress were made. At low stresses, the mate-
fluid to move through the system at 300 ml/min. rial had relatively constant and high apparent viscosity (104 Pa s).
The yield point was taken as the stress at which the viscosity
2.3. Continuous high pressure homogenization rapidly began to decline.
Measurements of shear rate dependency were made using a
The pasteurized puree was divided into four lots (2.5 L each) conical rotor 27.98 mm in diameter and 42.06 mm in length. In this
and processed in a Model nG7900 Ultra High Pressure case, the sample cup was filled with 30 ml of sample. The sample
Homogenizer (Model nG7900, Stansted Fluid Power Ltd.) at 10, was sheared at rates between 0.1 and 100 s1 over a period of
20, 30 and 40  103 PSI (69, 138, 207 and 276 MPa). Typical fluctu- 15 min. Plots of shear stress versus shear rate were developed
ations about the set-point were no greater than ±10 MPa. The pro- and fitted with a variety of models. Best fits were generally
duct was pumped to the systems intensifiers, raising the fluid to attained with a power law model:
the target pressure. The fluid was then directed to a stainless steel
micrometering valve (Model 60vrmm4882, Autoclave Engineers,
r ¼ K c_ n ð1Þ
Erie, PA). Samples were processed at the specified pressure for where K a consistency index, and n the power law index.
either 1, 2 or 3 passes. Between each run, or between cycles, the Viscoelastic properties were assessed using frequency-
product was held at 7 °C. Each treatment group had at least dependent small-strain oscillatory tests. Samples were tested
400 ml to be used for subsequent analyses. Processed samples using a parallel plate system 20 mm in diameter. The bottom plate
were kept at 2 °C in 500 ml Nalgene centrifuge bottles. was loaded with 250 ll of sample and the top plate brought to a
1 mm gap. Dynamic strains were set to 1% and the samples mea-
2.4. Moisture content sured in the range of 0.1–100 rad s1.

Moisture content of the puree, before and after processing, was 2.8. Color
determined using an automated moisture analyzer (Model HR73,
Mettler Toledo, Columbus, OH)). Samples (1 g) were loaded onto Color was measured with a Model CR-410 chroma meter
tared aluminum pans and weighed in the instrument. Moisture (Konica Minolta Sensing Americas, Ramsey NJ) with D65 illumi-
content was calculated based on mass before and after drying at nant and 10° CIE standard observer angle. Before the tests, the
95 °C in the halogen lamp system. As no subsequent evaporation chroma meter was calibrated with a standard white tile provided
was used, the solids in the processed samples ranged from 4.6 to the manufacturer. Samples were poured into 100  15 mm VWR
5.0 g/100 g. petri dishes to cover all areas to a thickness of 5 mm. All the petri
dishes were place on white print paper when tested. The data
2.5. Particle size analysis collected by the chroma meter was displayed in L⁄ (lightness),
a⁄(red–green axes) and b⁄(yellow–blue axes) values.
The particle size distribution was determined by laser diffrac-
tion using a Malvern Mastersizer (Model MSS, Malvern
2.9. Statistical methods
Instruments Ltd.). Samples were introduced into a stirred tank
filled with water until an obscuration of 20% was reached.
All tests were repeated at least three times and results and the
Between different samples, the detector and laser were aligned
means and standard deviations calculated. Measurements were
and backgrounds were calibrated. The presentation was selected
compared by two-way ANOVA using SAS 9.3 (SAS Institute Inc.,
as ‘‘Standard-Wet’’ (nH2O = 1.333, nparticle = 1.5295). Particle size
Cary NC). Factors included pressure (with levels of 69, 138, 207
was displayed in terms of the volume mean diameter (d4,3).
or 276 MPa) and numbers of passes (1, 2 or 3). Post hoc tests for
Cumulative percentiles d0.1, d0.5, and d0.9 were also determined,
significant differences among treatments was done using Tukey’s
indicating that 10%, 50% or 90% of the particles fell below the spec-
HSB test. The level of significance was set at p 6 0.05.
ified diameter.

2.6. Light microscopy 3. Results and discussion

Microstructural features of the processed purees were assessed 3.1. Moisture content
using light microscopy (Model DM100S, Leica Microsystems Inc.,
Buffalo Grove IL). Samples were diluted 1:10 with distilled water, Moisture content of the tomato purees after different high pres-
then 250 ll was placed on a microscope slide and mixed with sure processing regimens is shown in Table 1. Neither pressure
125 ll of 0.01 M methylene blue. Most samples were studied with level nor number of homogenization cycles resulted in differences
a 10 objective and 10 eyepiece. Digital images were attained in moisture content amongst treated samples, or compared with
with an 8 MP digital camera, and analyzed using the Spot the unprocessed control. The average moisture content of pro-
Imaging software (Sterling Heights, MI). cessed samples was 95.2 gH2O/100 g. There was some concern that
heat produced by pressure and shear effects might cause flashing
2.7. Rheological properties of moisture as the product exited the homogenizer. However, the
samples were rapidly cooled to 7 °C after each homogenization
Several rheological measurements were made on the processed step to limit evaporation and heat induced degradation of lycopene
tomatoes using a hybrid rheometer (Model Discovery HR-2, TA and other desirable constituents.
Instrument Inc., New Castle DE) at temperatures of 20 ± 0.1 °C.
Yield stress was determined using a static vane method. Tomato 3.2. Particle size distribution
puree samples (40 ml) were placed within the 33.7 mm diameter
sample cup. A four-bladed vane (28 mm diameter  42 mm long) Examples of the particle size distribution (PSD) following
was carefully lowered into the sample. The instrument was oper- high-pressure homogenization are shown in Fig. 1. Derived mea-
ated under stress-control mode, with the stress gradually surements of average particle diameter (d4,3) and cumulative per-
increased from 0.0 to 5.0 Pa over 10 min. Plots of apparent centiles (d0.1, d0.5, d0.9) are shown in Table 2. Both pressure level
48 J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54

Table 1 Table 2
Moisture content and color values for tomato puree processed at 69–276 MPa for 1–3 Mean and cumulative diameter of tomato puree as a function of CHP homogenization
passes. pressure and number of passes.

Pressure Number Moisture content L⁄ a⁄ b⁄ Pressure No. d4,3 d2,3 d0.1 d0.5 d0.9
(MPa) passes (%) (MPa) passes (lm) (lm) (lm) (lm) (lm)
Control 95.44a 56.29b 32.26b 35.56a Control 318h 216h 164i 311i 488j
69 1 95.26a 52.82a 37.01a 35.31a 69 1 153a 51.2a 31.0a 113a 337a
2 95.39a 52.36a 37.84a 35.05a 2 91.2b 26.5b 21.8b 70.5b 190b
3 95.26a 51.97a 37.91a 34.33a 3 81.2c 26.1b 21.1b 65.5c 164c
138 1 95.24a 52.84a 37.52a 35.67a 138 1 83.0c 23.9c 19.2c 63.5d 175bc
2 95.31a 52.82a 37.67a 35.55a 2 56.5e 17.5de 14.2d 46.5g 112ef
3 95.06a 52.28a 37.99a 35.52a 3 50.41ef 16.4ef 13.1e 42.7h 98.1fg
207 1 95.19a 52.59a 37.96a 35.39a 207 1 66.5d 18.4d 14.9d 51.8e 138d
2 95.18a 52.74a 38.09a 35.39a 2 46.2f 14.8f 12.6e 39.5i 88.9gh
3 95.10a 53.17a 37.64a 35.42a 3 34.3g 10.7g 9.05g 30.4j 65.0i
276 1 95.19a 52.82a 37.70a 34.98a 276 1 64.0d 17.7de 14.3d 49.7f 133de
2 95.00a 53.45a 37.07a 36.23a 2 36.7g 11.2g 10.4f 32.1j 69.6hi
3 95.01a 52.69a 37.93a 35.03a 3 31.1g 9.7g 7.82h 27.3k 60.6i
a
L⁄ – lightness on a scale from 0 (black) to 100 (white); a⁄ – values on the color axes Values in a column not followed by the same letter are significantly different at
ranging from green (100) to red (100); b⁄ – values on the color axes ranging from p < 0.05.
blue (100) to yellow (100).
a
Values in a column not followed by the same letter are significantly different at
p < 0.05.
In the CHP system, many factors influence the change in particle
size. A needle-seat micrometering valve was used in which the
pressure is determined by the force exerted on the needle blocking
and number of homogenization passes had a significant effect on the flow pathway. This creates tremendous shear stress as particles
particle size, as did the interaction between factors. As seen in speed through narrow pathways. In addition, particles impinge on
Fig. 1a, increasing pressure from 69 to 276 MPa decreased the aver- hard surfaces at relatively high velocities. Rapid changes in pres-
age mean diameter. Thus, after a single pass the mean particle sure may also create cavitation, creating bubbles in the fluid.
diameter had decreased from 318 lm to 153 lm at 69 MPa, to Fluctuating collapse of these voids create shock waves that may
83.0 lm at 138 MPa, to 66.5 lm at 207 MPa and to 64.0 lm at further disrupt particles. Finally, the high pressure and shear con-
276 MPa. There was no further reduction of particle size at pres- ditions cause heating of the fluid, which may further promote
sures higher than 207 MPa. Higher processing pressures also breakdown of particles. The diminishing effects of increased pres-
resulted in a narrower PSD. sure may be due to several reasons. First, it takes more than pro-
portionally higher pressures to force the suspension through a
narrower gap. Also, it is likely easier to disrupt clusters of many
tomato cells than to break apart small clusters and individual cells.
In addition, at sizes less than 250 lm cell wall fragments domi-
nate, and these are inherently more difficult to break apart than
adjacent cells.
Reduction in particle size with increasing pressure has also
been reported for other tomato products, although at lower
operating pressures. Augusto et al. (2012) investigated the
high-pressure homogenization of tomato juice at pressures up to
150 MPa. They found that unhomogenized control (4.5° Brix juice)
had particles in the size range 100–1000 lm. At 150 MPa, the par-
ticle size range was reduced to 10–300 lm. They also observed a
plateau at which increasing pressure resulted in only small further
changes in particle size. Homogenization pressures up to 9 MPa
decreased the particle size of ketchup (Bayod and Tornberg,
2011). For unprocessed paste, 50% of the particles were larger
than 100 lm in diameter (d3,2). After homogenization, a multi-
modal distribution of particle sizes was observed, and only 20%
of the particles were larger than 100 lm. Mert (2012) used
microfludization at up to 200 MPa to affect the properties of
ketchup. With a conventional valve homogenizer, they measured
d3,4 values of 328 lm, while microfluidization reduced d3,4 to
between 81 and 29 lm, for pressures between 120 and 200 MPa.
Similar effects on particle size have been noted for high pressure
processing of carrot dispersion, apple and citrus juice
(Lopez-Sanchez et al., 2011b; Donsì et al., 2009; Lacroix et al.,
2005).
Subsequent passes through the high-pressure homogenizer, at
the same pressure, resulted in even further reduction in particle
size (Fig. 1b). This is reflected in changes in the cumulative diam-
eters in Table 2. Thus, for control 10% of the particles had diameters
Fig. 1. Particle size distribution of tomato puree processed by continuous high-
pressure homogenization (CHP) (a) after 2 passes at 69–276 MPa, (b) after 1–3 less than 164 lm, while 90% had diameters less than 488 lm.
passes at 207 MPa. However at 207 MPa, for example, 10% had diameters less than
J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54 49

14.9 lm and 90% had diameters less than 138 lm. Hence, there number of cycles was a significant factor for K in the CHP-treated
was a narrowing of the range (d0.9–d0.1) from 324 lm to 123 lm. samples. After three passes in the homogenizer, the samples were
Most of the subsequent reduction in size was attained on the sec- always more viscous than after one or two passes. Amongst the
ond pass, with only a modest reduction occurring with the third CHP samples, those processed at the highest pressure had the low-
pass. For example, d4,3 decreased from 318 to 64.0 lm after one est K value after 1 cycle (538 mPa s), and the highest K value after
pass at 276 MPa, to 36.7 lm after two passes, and to 31.1 lm after 3 cycles (902 mPa s). Augusto et al. (2012) found viscosity in the
the third pass. The greater number of passes also led to narrower range of 500–900 mPa.s for tomato juice processed for one pass
PSD. As both pressure and number of passes were relevant factors, at up to 150 MPa. However, in their case viscosity tended to be
similar results might be attained by trading greater number of greater for samples processed at higher pressure, as did Mert
passes for higher pressure. Thus, processing tomato puree at (2012). For conventionally homogenized ketchup, the zero-shear
138 MPa resulted in d4,3 of 56.5 lm after 2 passes, slightly smaller viscosity was on the order of 20,000–50,000 mPa s. For tomato
than a d4,3 of 64.0 lm realized after 1 pass at 276 MPa. paste, the viscosity depended on solids concentration, ranging
from as low as 3000 mPa s to as high as 600,000 mPa s.
3.3. Microstructure The power law index n was least for the control puree (0.12),
indicating greater shear-thinning behavior than for the
Fig. 2 shows microphotographs of the tomato puree processed CHP-treated samples. While n was greater for all CHP samples
at different pressures, and number of passes through the CHP (0.28–0.39), there was no particular trend associated with pressure
homogenizer. Most of the particulate structure of the puree con- or number of passes. Thixotropic properties generally occur in flu-
sists of cell fragments, and some intact cells (Bayod et al., 2007). ids with some degree of continuous microstructure, which is dis-
Most of these are relatively soft parenchymal cells associated with rupted by a finite stress and subsequent flow field. The
the mesocarp, as much of the skin and seeds were removed in pro- associations may be due to flocculation, alignment of fibers, entan-
cessing. In general, changes in the microstructural features of the glements and molecular associations. Shear-thinning occurs in the
puree mirrored the results of particle size analyses. Generally, high viscous flow region due to alignment of elongated particles, disrup-
homogenization pressures caused a reduction in particle size. Most tion of junction zones and break-up of flocculated particles
notably, larger elongated tissue fragments were reduced to pieces (Barnes, 1997). For the CHP-treated samples, the particles are
with more uniform dimensions in all directions. At lower pressures smaller and likely more spherical in nature, factors which generally
(69 and 138 MPa), particles were reduced in size as compared to decrease thixotropy.
control, but there were still evidence of stained clusters up to As noted by Augusto et al. (2011), tomato products are a
337 lm in length. At greater pressure levels (207 and 276 MPa), two-phase system with complicated rheology. The aqueous serum
much of these clusters were reduced in size. In addition, it became phase contains sugars, salts, acids and some soluble polysaccha-
more difficult to see distinct outlines of the cellular material, sug- rides, but is by and large a Newtonian fluid. The dispersed phase
gesting that the cells became more disrupted and the stained contains clusters of cells, single cells, fragmented cells or cell rem-
polysaccharides were less restricted. This suggests at very high nants, as well as colloidally dispersed polysaccharides. The system
pressure, more of the cellular array was disrupted, releasing the is best thought of as a non-colloidal dispersion, as many of the par-
cell wall material into the surrounding medium. ticles are above 10 lm in size. As such, the flow properties of the
Increasing the number of passes through the CHP system also material are dominated by hydrodynamic forces, and less by inter-
helped reduce overall particle size. In addition, greater number of particle forces and Brownian motion (Genovese et al., 2007). Other
passes created a more uniform distribution of particle sizes. factors such as particle size distribution and shape are also impor-
Greater pressure and number of passes also created particles that tant in this regime. Of course, fruit and vegetable juice and purees
were both stained and had areas of translucency. As the methylene may contain particles in the noncollodial range (>10 lm) as well as
blue stains acidic regions of the cell, and particularly nuclear mate- the colloidal range (1–10 lm). For the former case, the simplest
rial, the appearance of more transparent matter indicates cell rup- description of the relative viscosity was postulated by Einstein as
ture and the release of cytoplasmic or cell wall material. In (Hiemenz and Rajagopalan, 1997):
addition, for purees processed at 207 or 276 MPa, there was a
greater tendency for particles to re-associate, presumably due to
g
gr ¼ ¼ 1 þ ½g/ ð2Þ
the released pectin and cell wall materials. Similar results have gs
been reported for tomato juice by Augusto et al. (2012) and Kudo
where gs is the viscosity of the solvent and g the viscosity of the dis-
et al. (2013), as well as for ketchup by Mert (2012).
persion. According to this viewpoint, the viscosity increases as a
function of /, the relative volume fraction of particles dispersed
3.4. Rheological properties
in the total volume. The Einstein equation holds primarily in dilute
suspensions, however. For concentrated dispersions, the Krieger–
Fig. 3 shows a plot of shear stress versus shear rate for the
Dougherty equation is often applied:
unhomogenized and CHP treated samples. Data between 1 and
45 s1 were fit to a power law equation, and fitted parameters K  ½g/m
/
and n are summarized in Table 3. At shear rates below 1 s1 the gr ¼ 1  ð3Þ
/m
samples showed evidence of elastic behavior. That is, there was
an initial precipitous increase in shear stress, followed by a where /m is the maximum packing fraction of solids, and typi-
decrease. As this was evidence of a static yield stress, this was cally varies from 0.60 to 0.74, depending on the type of packing
examined by a different method. Interestingly, including a yield (Genovese et al., 2007). It can be determined experimentally
stress term (ro) in the power law model did not improve the fits. through sedimentation experiments. This equation may be modi-
Thus, once the elastic network was broken and flow was initiated, fied further to account for various particle shapes and deformabil-
the apparent viscosity was relatively low. ity, as well as interaction amongst particles.
The consistency index (K) can be taken as the apparent viscosity Both microscopic and particle size analyses showed that the
at low-shear rate (1 s1). The control had the greatest K suspended tomato particles were reduced in size with higher pres-
(3474 mPa s), and all CHP homogenized samples had significantly sure, or greater number of passes through the CHP system.
lower values (538–902 mPa s). Two-way ANOVA showed that the Concomitantly, the viscosity at low shear rates was reduced. Two
50 J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54

Fig. 2. Photomicrographs of tomato puree after CHP processing. From top to bottom: increasing pressure from 69 MPa to 276 MPa; from left to right: increasing number of
passes (1–3).
J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54 51

in the range 0.63–1.0 mm were less shear-thinning than those with


particles in the range 0.35–0.63 mm or <0.35 mm. Lopez-Sanchez
et al. (2011a) studied emulsions made from vegetable oil, and
either carrot or tomato, and subject to homogenization pressures
of 10 or 100 MPa. They found that samples processed at the higher
pressure were more shear-thinning than those processed at lower
pressure. In our case, processed samples were less shear-thinning
than unprocessed pulp, but there were no clear differences
amongst samples processed at different pressures and number of
passes. Any differences between this study and other researches
may be due to the greater homogenization pressures used here,
different microstructural features (e.g. dispersed pulp versus emul-
sions), and lower particle sizes attained in this study.
The results from yield stress measurements are shown in
Table 4. The control puree had the highest yield stress at 2.62 Pa.
All treated samples had lower yield stress, although the differences
Fig. 3. Shear stress versus shear rate for control () and CHP samples processed two were not large. After one pass, samples processed at 276 MPa had
times at (N) 69, (}) 138, (s) 207 and (j) 276 MPa. the lowest yield stress (1.09 Pa), those at 138 or 207 MPa slightly
higher yield stress (1.70–1.85 Pa), and those at 69 MPa even higher
factors may be at play. First, in addition to volume fraction the yield stress (2.26 Pa). In several cases, however, increasing the
actual particle size and distribution also determines the viscosity. number of passes in the CHP system led to greater yield stress than
While Equations (2) and (3) do not implicitly involve particle size, found after 1 pass. Thus at 276 MPa, the yield stress was 1.09 Pa
particle size is known to influence the relative viscosity in some after one pass, 1.69 Pa after two passes, and 2.12 Pa after three
cases. For example, Pavlik (2011) measured the viscosity of disper- passes.
sions with glass particles ranging from 26 to 71 lm in diameter. Although the samples did have a measureable yield stress, they
The viscosity of dispersions containing particles between 36 and were not large compared to some other food products. For exam-
71 lm varied only slightly with / even at relatively high volume ple, pancake batter has a yield stress of 10 Pa, ketchup 15 Pa, and
fractions. Dispersions with 26 lm particles, however, had dis- mayonnaise 100 Pa. Abu-Jayil et al. (2004) studied the yield stress
tinctly lower viscosity, particularly at high volume fractions. of tomato paste with solids content ranging from 5.57% to 25% by
They believed that suspensions with smaller particles generated weight, and found the yield stress ranged from 1.6 to 101.9 Pa.
more open space between neighboring particles, allowing the con- Mert (2012) found that the yield stress of ketchup ranged from
tinuous phase to flow, and producing a fluid that behaved more 15.1 to 20.9 Pa for samples processed by microfluidization at pres-
like a dilute suspension that could be modeled by Einstein’s equa- sures between 20 and 200 MPa. For samples processed by a con-
tion. There may also be differences in the maximum packing frac- ventional valve homogenizer, the yield stress was only 9.4 Pa.
tion /m due to changes in particle size and distribution, shape or This highlights that the type of particle size reduction is important
even deformability. In addition, the CHP processing may also to the final product properties. With microfluidization, two fluid
reduce the actual volume fraction of particles, particularly if cells microstreams collide at high speed. The streams are subjected to
are disrupted and disperse their contents, creating suspensions high shear and substantial impact forces. With the valve homoge-
that are more appropriately characterized as colloidal. nizer, the fluid is forced through an orifice created by a valve and
Some studies have reported an increase in shear-thinning prop- valve seat. This creates high turbulence and shear, as well as cavi-
erties after processing of fruit and vegetable products. Silva et al. tation and impact.
(2010) found an increase in shear-thinning for pineapple pulp pro- The role of dispersed phase properties on yield stress (ro) is not
cessed at up to 30 MPa. When processed at greater than 40 MPa, completely understood. In fact, the term yield stress may be just a
the pulp exhibited phase separation due to increased aggregation. practical designation. Certainly, some viscoelastic materials behave
Shijvens et al. (1998) used screening procedures to produce apple- elastically at stresses below the yield value (r < ro), and flow at
sauce with various size fractions of particles. Those with particles higher stresses (r > ro). However, most of these would likely yield

Table 3
Rheological parameters for CHP processed tomato puree at 1–45 s1 as fit by the Table 4
powerlaw model (Eq. (1)). Yield stress for tomato puree CHP processed at 69–276 MPa for 1–3 passes.

Pressure (MPa) Number passes K (s) (mPa sn) Rate index n Pressure (MPa) Number passes Yield stress (Pa)
a c
Control 3474 0.12 Control 2.62e
69 1 771bc 0.28ab 69 1 2.26a b
2 652cd 0.37ab 2 2.07b
3 750bc 0.34ab 3 2.10b
138 1 623cd 0.39ab 138 1 1.85c
2 649cd 0.36a 2 2.14b
3 868b 0.34b 3 2.27a b
207 1 584cd 0.39a 207 1 1.70c
2 869b 0.36ab 2 2.12b
3 878b 0.36ab 3 2.45a
276 1 538d 0.41a 276 1 1.09d
2 771bc 0.38ab 2 1.69c
3 902b 0.36ab 3 2.12b
a a
Values in a column not followed by the same letter are significantly different at Values in a column not followed by the same letter are significantly different at
p < 0.05. p < 0.05.
52 J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54

stress of applesauce was greatest for particle diameter of 0.5 mm


and 100% pulp content, and least for 1.0 mm particles and 80% pulp
content. This may be an indication that increased surface area
enhanced interparticle attractive forces. Espinosa et al. (2011)
found that yield stress of apple puree increased with pulp content
and decreased with decreasing particle size. For tomato pulp, Yoo
and Rao (1994) showed that yield stress increased with solids con-
centration, but was generally greater for purees with particles sizes
0.34 mm as compared to those 0.71 mm. However, at relatively
low concentrations (11–12% by weight) purees with smaller parti-
cles had lower or no different yield stress than those with larger par-
ticles. Bayod et al. (2007) observed that homogenized tomato paste
had greater yield stress than non-homogenized paste. They attribu-
ted this to a change in the nature of the network structure.
Homogenized samples have fewer intact cells and cell clusters, and
are more fibrous in nature. In addition, homogenization creates more
surface area that may enhance interparticle forces of attraction.
Fig. 4 and Table 5 show how the complex shear moduli vary
with frequency. In all cases, the storage modulus G0 was greater
than the loss modulus G00 . Thus, at 1% strain (r  0.16 Pa), the sys-
tem behaves as an elastic solid, and might be characterized as a
soft gel network (Rao and Cooley, 1992). For the unprocessed con-
trol, the G0 at low frequency (0.1 rad/s) was 148.7 PA, while that for
all processed samples ranged from 10.9 to 21.0 Pa. A measurable
value of G0 in the viscous region is typical for dispersions and con-
centrated polymer solutions. It has been suggested that this is due
to the relative dense packing of particles, interparticle forces, and
entanglements of cells, fibers, and large colloidal molecules
(Mert, 2012). These are the same factors, of course, that contribute
to yield stress. It is worth noting that G0 at both 0.1 and 1.0 rad/s
were well correlated and decreased with particle size (d4,3). That is:

x ¼ 0:1rads1 ! G0 ¼ 0:0023  d24;3  0:315  d4;3 þ 25:9 ðr2 ¼ 0:993Þ


ð4Þ
Fig. 4. Storage (G0 ) and loss modulus (G00 ) (a) as a function of frequency for tomato
puree processed at 207 MPa for 1–3 passes; (b) puree processed in 1 pass at 69– x ¼ 1:0rads1 ! G0 ¼ 0:0031  d24;3  0:486  d4;3 þ 37:1 ðr2 ¼ 0:993Þ
276 MPa.
ð5Þ
at r < ro given sufficient time. However, Bayod et al. (2007) All samples also became notably stiffer at x > 3 rad/s. Below
showed that tomato products deformed at r < ro maintained elas- 3 rad/s, G0 / x0.2; at x > 3 rad/s, G0 / x1.8. True gels are more
tic behavior for up to 30 min. In general, yield stress is greater for likely to have chemical bonds that hold the network, and show lit-
systems with greater volume fraction of particles and stronger tle frequency dependence. The frequency dependency shown by
interaction forces amongst the particles (Genovese et al., 2007). the tomato samples is more typical of ‘‘weak gels’’ that are held
As noted, yield stress values increased with solids concentration, together by physical associations. At low frequencies, some of the
and therefore /, in tomato pastes (Mert, 2012). structure behaves elastically, but there are still relaxation or flow
The influence of particle size on yield stress is less well under- processes that can occur. At higher frequencies, the time scale is
stood, contradictory in some cases, and depends on the specific too short for these elements to flow, and the structure as a whole
particle range of interest. Qui and Rao (1988) reported that yield behaves more elastically. Similar behavior has been observed for

Table 5
Storage (G0 ) and loss (G00 ) moduli for CHP treated tomato puree.

Pressure (MPa) No. passes G0 at 0.1 s1 (Pa) G0 at 1 s1 rad/s (Pa) G00 at 0.1 s1 (Pa) G00 at 1 s1 (Pa) tan (d) tan (d)
Control 148.74g 199.52g 46.03e 30.21g 0.31a 0.15a
69 1 20.98a 27.97a 3.48a b 2.59a 0.17c 0.092b
2 17.12b 23.20bc 2.80bc 2.19a bc 0.17c 0.091b
3 16.86b 23.26bc 2.85bc 2.17a bc 0.17c 0.094b
138 1 16.06bc 22.03cd 2.75bc 2.00cd 0.17c 0.096b
2 16.82b 23.11bc 2.85bc 2.11bcd 0.16c 0.094b
3 15.98bc 22.33bcd 2.87bc 1.90cde 0.17c 0.091b
207 1 12.79ef 17.76ef 2.13cd 1.66de 0.18c 0.092b
2 15.36bcd 21.52cde 2.81bc 1.99cd 0.18c 0.094b
3 17.08b 26.12a b 3.86a 2.54a b 0.17c 0.093b
276 1 10.94f 15.39f 1.87d 1.48e 0.18c 0.085b
2 13.34def 18.94def 2.42cd 1.77cde 0.22b 0.097b
3 14.10cde 19.90cde 2.60cd 1.80cde 0.18c 0.091b
a
Values in a column not followed by the same letter are significantly different at p < 0.05.
J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54 53

tomato homogenates subject to static high pressure (Verlent et al., size distribution. The average particle size was reduced by up to
2006). However, tomato juice showed little viscoelastic behavior, ten-fold, with sizes on the order of 30–80 lm made possible.
with G0 varying little with frequency over the range 0.01–10 rad/s. Although not studied in this work, this is in the size range in which
For tomato paste, Sanchez et al. (2002) found that G0 was depen- consumers no longer perceive a particulate structure. These
dent on x0.12 over the range of 0.01–100 rad/s. changes were made possible in a continuous manner, and would
G0 values shown in Table 5 are in general lower than those reported not be realized with static high pressure alone. The change in size
by other researchers working on related products. Bayod et al. (2008) and shape of particles in the dispersed phase also altered the rhe-
studied tomato paste and ketchup with different levels of total solids ological properties of the puree. The material was less viscous at
(23.42–38.75% TS). For tomato paste, G0 at 1 Hz ranged from 7124 to low shear and with a lower yield stress, but had less shear thinning
10,411 Pa while for ketchup G0 ranged from 562 to 732 Pa. In our behavior. This has implications for developing products with
study, however, the total solids level was much lower (5%), as were greater solids, as well as for pumping requirements during pro-
the size of particles. For microfluidized tomato ketchup, Mert (2012) cessing. This may also make it easier to produce spray-dried and
reported an increase in G0 for pressures between 20 and 120 MPa, similar products.
and a decrease for pressures from 120 to 200 MPa.
For samples processed at 69 MPa, additional passes through the References
CHP system led to a lower G0 , while at 207 or 276 MPa it led to a
Abu-Jdayil, B., Banat, F., Jumah, R., Al-Asheh, S., Hammad, S., 2004. A comparative
greater G0 . However, at 69 MPa, additional passes reduced the par- study of rheological characteristics of tomato paste and tomato powder
ticle size (d4,3) of relatively big particles from 153 to 91 and 83 lm. solutions. Int. J. Food Prop. 7 (3), 483–497.
At 276 MPa, after 1 pass the particle size was already 64 lm; addi- Augusto, P.E., Falguera, V., Cristianini, M., Ibarz, A., 2011. Viscoelastic properties of
tomato juice. Proc. Food Sci. 1, 589–593.
tional passes reduced it to 37 and 31 lm. One theory is that in the Augusto, P.E., Ibarz, A., Cristianini, M., 2012. Effect of high pressure homogenization
former case reducing particle size was more important to reducing (HPH) on the rheological properties of tomato juice: time-dependent and
G0 by allowing these large particles to better slip pass each other steady-state shear. J. Food Eng. 111 (4), 570–579.
Barnes, H.A., 1997. Thixotropy-a review. J. Non-Newtonian Fluid Mech. 70, 1–33.
when under stress. In the latter case, the further reduction of rela- Bayod, E., Tornberg, E., 2011. Microstructure of highly concentrated tomato
tively small particles allowed more surface area and interaction suspensions on homogenisation and subsequent shearing. Food Res. Int. 44
amongst particles, but in this case colloidal forces rather than hydro- (3), 755–764.
Bayod, E., Månsson, P., Innings, F., Bergenståhl, B., Tornberg, E., 2007. Low shear
dynamic forces are more dominant. Another consideration is that rheology of concentrated tomato products. Effect of particle size and time. Food
subsequent homogenizer passes allow greater swelling of the cell Biophys. 2 (4), 146–157.
fragments, increasing their effective phase volume in the suspension Bayod, E., Willers, E.P., Tornberg, E., 2008. Rheological and structural
characterization of tomato paste and its influence on the quality of ketchup.
(Lopez-Sanchez et al., 2011b; Bayod and Tornberg, 2011).
LWT – Food Sci. Technol. 41 (7), 1289–1300.
Table 5 also lists tan(d) at 1 rad/s, which is the ratio of loss to Colle, I., van Buggenhout, S., van Loey, A., Hendrickx, M., 2010. High pressure
storage modulus (G00 /G0 ). For the unprocessed control, tan(d) = 0.31. homogenization followed by thermal processing of tomato pulp: influence on
For the CHP treated samples, tan(d) was smaller with values rang- microstructure and lycopene in vitro bioaccessibility. Food Res. Int. 43 (8), 2193–2200.
Colle, I.J., Andrys, A., Grundelius, A., Lemmens, L., Lofgren, A., van Buggenhout, S.,
ing from 0.16 to 0.18 (with the exception of tan(d) = 0.22 for sam- van Loey, A., Hendrickx, M., 2011. Effect of pilot-scale aseptic processing on
ples treated for 2 passes at 276 MPa). Similar results have been tomato soup quality parameters. J. Food Sci. 76 (5), C714–C723.
reported for tomato juice (Lopez-Sanchez et al., 2011a) and other Corbo, M.R., Bevilacqua, A., Campaniello, D., Ciccarone, C., Sinigaglia, M., 2010. Use
of high pressure homogenization as a mean to control the growth of foodborne
plant products (Pickardt et al., 2004; Romana and Taylor, 1992). moulds in tomato juice. Food Contr. 21 (11), 1507–1511.
This means that while the control had higher values of G0 , the vis- Donsì, F., Esposito, L., Lenza, E., Senatore, B., Ferrari, G., 2009. Production of shelf-
cous component G00 was proportionally greater than for the pro- stable annurca apple juice with pulp by high pressure homogenization. Int. J.
Food Eng. 5 (4), 1556–3758.
cessed samples. One postulate is that while smaller particles and Espinosa, L., To, N., Symoneaux, R., Renard, C.M., Biau, N., Cuvelier, G., 2011. Effect of
less entanglements led to a reduced storage modulus, it also processing on rheological, structural and sensory properties of apple puree.
resulted in a lower viscosity of the lossy component. Proc. Food Sci. 1, 513–520.
Etminan, M., Takkouche, B., Caamaño-Isorna, F., 2004. The role of tomato products
and lycopene in the prevention of prostate cancer: a meta-analysis of
3.5. Color observational studies. Cancer Epidemiol., Biomarkers Prevent. 13 (3), 340–345.
Food and Agriculture Organization (FAO), 2012. Production of Tomato by countries.
Measured color parameters are shown in Table 1 for the differ- http://en.wikipedia.org/wiki/Tomato, 2014-09-11/2014-10-01 [Retrieved
21.02.14].
ently processed tomato purees. Lightness value (L⁄) for the control Floury, J., Desrumaux, A., Legrand, J., 2002. Effect of ultra-high-pressure
was 56.3. CHP processed samples were somewhat darker homogenization on structure and on rheological properties of soy protein-
(L⁄ = 51.9–53.2), but there were no differences in lightness stabilized emulsions. J. Food Sci. 67 (9), 3388–3395.
Garcia, A.F., Butz, P., Tauscher, B., 2001. Effects of high-pressure processing on
amongst the CHP treated samples. While there were no differences
carotenoid extractability, antioxidant activity, glucose diffusion, and water
in b⁄ amongst any samples, the control had slightly lower a⁄ (32.3) binding of tomato puree (Lycopersicon esculentum Mill.). J. Food Sci. 66 (7),
as compared to any of the CHP treated samples (a⁄ = 37.1–38.1). 1033–1038.
Thus, CHP treated samples were slightly more red in color. A pre- Genovese, D.B., Lozano, J.E., Rao, M.A., 2007. The rheology of colloidal and
noncolloidal food dispersions. J. Food Sci. 72 (2), R11–R20.
vious study by Rodrigo et al. (2007) showed that no color degrada- Giovannucci, E., 1999. Tomatoes, tomato-based products, lycopene, and cancer:
tion of tomato puree occurred during combined thermal and high review of the epidemiologic literature. J. National Cancer Inst. 91 (4), 317–331.
hydrostatic pressure treatments. This implies that lycopene, the Herh, P.K., Colo, S.M., Roye, N., Hedman, K., 2000. Rheology of foods: new
techniques, capabilities, and instruments. Am. Labor. 32 (12), 16–21.
most abundant carotenoid in tomato and responsible for the red Hernádez, A., Cano, M.P., 1998. High-pressure and temperature effects on enzyme
color (Colle et al., 2011), was not lost during processing. inactivation in tomato puree. J. Agric. Food Chem. 46 (1), 266–270.
However, some studies have shown that high-pressure treatments Hiemenz, P.C., Rajagopalan, 1997. Ch. 4: The rheology of dispersions. In: Principles
of Colloid and Surface Chemistry. Marcel Dekker, Inc., New York.
might reduce the storage stability of lycopene. For example, Kudo Kirsh, V.A., Mayne, S.T., Peters, U., Chatterjee, N., Leitzmann, M.F., Dixon, L.B., Urban,
et al. (2013) found that tomato juice treated at 75 and 100 MPa had D.A., Crowford, E.D., Hayes, R.B., 2006. A prospective study of lycopene and
significant color degradation after 60 days storage. tomato product intake and risk of prostate cancer. Cancer Epidemiol.,
Biomarkers Prevent. 15 (1), 92–98.
Kudo, M.T.K., Augusto, P.E., Cristianini, M., 2013. Effect of high pressure
4. Conclusions homogenization (HPH) on the physical stability of tomato juice. Food Res. Int.
51 (1), 170–179.
Lacroix, N., Fliss, I., Makhlouf, J., 2005. Inactivation of pectin methylesterase and
Continuous high pressure processing of tomato puree resulted stabilization of opalescence in orange juice by dynamic high pressure. Food Res.
in substantial changes in particle size and uniformity of particle Int. 38 (5), 569–576.
54 J. Tan, W.L. Kerr / Journal of Food Engineering 166 (2015) 45–54

Lopez-Sanchez, P., Svelander, C., Bialek, L., Schumm, S., Langton, M., 2011a. Rao, M.A., Cooley, H.J., 1992. Rheological behavior of tomato pastes in steady and
Rheology and microstructure of carrot and tomato emulsions as a result of dynamic shear. J. Texture Stud. 23 (4), 415–425.
high-pressure homogenization conditions. J. Food Sci. 76 (1), E130–E140. Rodrigo, D., van Loey, A., Hendrickx, M., 2007. Combined thermal and high pressure
Lopez-Sanchez, P., Nijsse, J., Blonk, H.C., Bialek, L., Schumm, S., Langton, M., 2011b. colour degradation of tomato puree and strawberry juice. J. Food Eng. 79 (2),
Effect of mechanical and thermal treatments on the microstructure and 553–560.
rheological properties of carrot, broccoli and tomato dispersions. J. Sci. Food Sanchez, M.C., Valencia, C., Gallegos, C., Ciruelos, A., Latorre, A., 2002. Influence of
Agric. 91 (2), 207–217. processing on the rheological properties of tomato paste. J. Sci. Food Agric. 82
McKenna, B.M., 2003. Introduction to food rheology and its measurement. In: (9), 990–997.
Mckenna, B.M., Lyng, J.G. (Eds.), Texture in Food. University College Dublin, Schijvens, E.P.H.M., Vliet, T.V., Dijk, C.V., 1998. Effect of processing conditions on the
Ireland, pp. 130–132. composition and rheological properties of applesauce. J. Texture Stud. 29 (2),
Mert, B., 2012. Using high pressure microfluidization to improve physical properties 123–143.
and lycopene content of ketchup type products. J. Food Eng. 109 (3), 579–587. Silva, V.M., Sato, A.C.K., Barbosa, G., Dacanal, G., Ciro-Velásquez, H.J., Cunha, R.L.,
Pavlik, M., 2011. The dependence of suspension viscosity on particle size, shear rate, 2010. The effect of homogenisation on the stability of pineapple pulp. Int. J.
and solvent viscosity. College of Liberal Arts & Social Sciences Theses and Food Sci. Technol. 45 (10), 2127–2133.
Dissertations. Paper 71. http://via.library.depaul.edu/etd/71, 2014-09-14/2014- Stang, M., Schuchmann, H., Schubert, H., 2001. Emulsification in high-pressure
10-21. homogenizers. J. Life Sci. Eng. 1 (4), 151–157.
Pickardt, C., Dongowski, G., Kunzek, H., 2004. The influence of mechanical and Tabilo-Munizaga, G., Barbosa-Cánovas, G.V., 2005. Rheology for the food industry. J.
enzymatic disintegration of carrots on the structure and properties of cell wall Food Eng. 67 (1), 147–156.
materials. Euro. Food Res. Technol. 219 (3), 229–239. Thakur, B.R., Singh, R.K., Handa, A.K., 1995. Effect of homogenization pressure on
Porretta, S., Birzi, A., Ghizzoni, C., Vicini, E., 1995. Effects of ultra-high hydrostatic consistency of tomato juice 1. J. Food Qual. 18 (5), 389–396.
pressure treatments on the quality of tomato juice. Food Chem. 52 (1), 35–41. Verlent, I., Hendrickx, M., Rovere, P., Moldenaers, P., Loey, A.V., 2006. Rheological
Qian, C., McClements, D.J., 2011. Formation of nanoemulsions stabilized by model properties of tomato-based products after thermal and high-pressure
food-grade emulsifiers using high-pressure homogenization: factors affecting treatment. J. Food Sci. 71 (3), S243–S248.
particle size. Food Hydrocolloids 25 (5), 1000–1008. Yoo, B., Rao, M.A., 1994. Effect of unimodal particle size and pulp content on
Qiu, C.G., Rao, M.A., 1988. Role of pulp content and particle size in yield stress of rheological properties of tomato puree. J. Texture Stud. 25 (4), 421–436.
apple sauce. J. Food Sci. 53 (4), 1165–1170. Yuan, Y., Gao, Y., Zhao, J., Mao, L., 2008. Characterization and stability evaluation of
Ramana, S.V., Taylor, A.J., 1992. Development of a method to measure rheological b-carotene nanoemulsions prepared by high pressure homogenization under
properties of carrot cell and cell-wall materials. J. Sci. Food Agric. 60 (1), 39–45. various emulsifying conditions. Food Res. Int. 41 (1), 61–68.

You might also like