You are on page 1of 5

Rheological Properties of Tomato Concentrates as

Affected by Particle Size and Methods of Concentration


T. TANGLERTPAIBUL (nee SORNSRIVICHAI) and M. A. RAO

ABSTRACT

Shear rate-shearstressdata were obtainedon tomato concentrates


madefrom juices that were producedusing finisher screenopenings
(FSO): 0.020, 0.027, 0.033, and 0.045 in. In general,the apparent
viscosity of the concentratesat a shearrate of 100 set- increased
with increasein FSO. However, concentratesmadefrom juice using
a 0.027 in FSO had the highest apparentviscosity. Magnitudesof
yield stressof concentratesincreasedin direct proportion to FSO.
Apparentviscositiesof concentrates
madeby evaporatingtomatojuice
were lower than thoseobtainedeither by evaporatingthe serumor by
reverseosmosisconcentrationof the serum.
INTRODUCTION

THERE IS a considerable volume of literature in which the


viscosity or consistency characteristics or other physical properties of tomato juice and concentrates have been related to
processsingconditions (Kertesz and Loconti, 1944; McColloch
et al., 1950; Davis et al., 1954; Whittenberger and Nutting,
1957, 19.58;Kopelmann and Mannheim, 1964; Mannheim and
Kopelmann, 1964). In most cases, however, the viscosities or
consistencies reported were the results of single-point measurements; that is, a single viscosity or consistometer time was
given for each sample. Few studies have taken the non-Newtonian nature of tomato concentrates into consideration.
The Bostwick consistometer values of tomato concentrates
are related to the insoluble solids by an exponential relationship
(Marsh et al., 1977) that become very small (<l cm) as the
concentration increases so that they cannot be determined for
solids concentration more than about 15%. Tomato juice characteristics might be expected to depend partly on the structure
of the original cells and cell walls (Hand et al., 1955). Kertesz
and Loconti (1944) found that the size and shape of suspended
cell wall particles reflected the severity of mechanical stresses
applied by finisher (or screen).
Particle size distribution is a major contributor to the viscosity of tomato juice (Surak et al., 1979). When tomato juice
is passed through a fine finishing screen, small particles are
incorporated in the juice and a high proportion of the particles
remain spherical. The diameter of the holes in the finisher
screen affects the particle size and the juice viscosity (Kattan
et al., 1956; Smit and Nortje, 1958). The speed (rpm) of operation of the paddle finisher also has an effect on juice viscosity (Whittenberger and Nutting, 1957). At higher operating
speedsthere is an increase in the number of particles produced
and the particles tend to be more elongated and the juice has
higher viscosity (Hand et al., 1955).
The quantity, configuration, and characteristics of the suspended particles also influence the viscosity of the product
(Tanford, 1961; Hand et al., 1955; Robinson et al., 1956; Luh
et al., 1954). However, Luh et al. (1956) found no correlation

Author Rae is with the Dept. of food Science & Technology,


Cornell Univ., N YS Agricultural Experiment Station, Geneva, NY
14456. Author Tanglertpaibul
(nee Sornsrivichai), formerly with
Cornell Univ., is with Mah Boonkrong Center, Bangkok, Thailand. Address inquiries to Dr. Rao.

between cellulose content, which is the major component in


cell walls, and the consistency of tomato juice. It appears that
not only the quantity of particles affects consistency but their
characteristics, such as shape and size have a more direct effect
on consistency.
Microscopic examination of samples indicated that tomato
products finished with a screen with large holes did not consist
of uniformly large particles but rather of a mixture of small
particles along with suspendedshreds of tissue, the amount of
the latter increasing with the screen size (Kattan et al., 1956).
Kertesz and Loconti (1944) reported that under the microscope
and with proper staining many of these suspended particles
exhibit tom, ragged shape.
The use of fine screens for the reduction of particle size
would enhance oxidation because of relatively large surface
exposed to air and metal. Therefore, increasing the screen size
increased the retention of color and ascorbic acid (Kattan et
al., 1956).
Harper and El Sahrigi (1965) presented a relationship between apparent viscosity at 500 set- , concentration and temperature for a sample obtained by direct evaporation of tomato
juice. The shear rate ranged from 500 to 800 set- and concentration ranged from 12.8 to 30% total solids. They also
removed the insoluble solids in tomato juice by centrifugation,
concentrated the serum by evaporation to 65 Brix, and reconstituted the components. Harper and El Sahrigi (1965) found
that the apparent viscosities of the reconstituted samples were
only about one-third of those of the corresponding original
concentrates. A similar result was observed by Mannheim and
Kopelmann ( 1964).
Rao et al. (1981) studied the rheology of tomato concentrates from four different cultivars and arrived at a viscosity
relationship simil.ar to that of Harper and El Sahrigi (1965).
For the cultivars studied, the magnitude of the concentration
exponent was found to be 2.4 to 2.6.
The yield stress refers to the stress that must be exerted to
just move one fluid layer past another (Charm, 1962). On this
basis the yield stress has been related to the strength of the
coherent network structure as the force per unit area required
to breakdown the structure, followed by a rupture of the network bonds or linkages connecting the flow units (Dzuy and
Boger, 1983).
Because of the small magnitudes of yield stress of many
foods, it is difficult to determine its value experimentally.
Therefore, it has been found convenient to fit the experimental
shear stress-shearrate data to one or more of the constitutive
equations proposed by Herschel and Bulkley (1926), Casson
(1959), and Mizrahi and Berk (1972). The application of these
models to tomato concentrateswas studied by Rao et al. (198 1)
and Rao and Cooley (1983).
The overall objective of the present investigation was to
study the rheological properties of hot break tomato juice and
concentrates that contained different sized pulp particles and
that were concentrated in different manners. For this purpose,
concentrates were prepared from juices that were produced
using finisher screenswith different hole diameters. Using juice
from one finisher screen, concentrates were also produced by
three methods: (1) by evaporation of tomato juice, (2) from
Volume 52, No. 1, 1987-JOURNAL

OF FOOD SCIENCE-141

EFFECT OF PROCESSING

ON TOMATO VISCOSITY. .

evaporated tomato serum, and (3) from serum that was concentrated by reverse osmosis.
MATERIALS
Preparation

& METHODS

of tomato juice

Natural

tomato soluble solids determination

An A0 ABBE Refractometer (American Optical Corp., Buffalo,


NY) was used to determine the natural tomato soluble solids (NTSS)
in Brix. In the case of a very concentrated samples, only its serum
portion was used for NTSS determination because the presence of the
pulp in large amounts obscured the reading.

One hundred sixty kilograms of fresh tomatoes, cultivar FM6203,


were hand-picked from commercial fields in Erie county, New York,
stored in a 21.1C room overnight and processed in the Experiment
Stations pilot plant. The tomatoes were sorted, washed, and crushed
in a hammer mill (W.J. Fitzpatrick Co., Chicago, IL) and the macerate was heated immediately in a steam-jacketed kettle to 98C. The
time interval between crushing the tomatoes and for the macerate to
reach 98C was approximately 1-2 min. The macerate was held at
98C for 4 min and 5 set as calculated using a z-value of 8.33C
(Nelson and TressleT, 1980), and using t, of 45 set at 103.9C (Bimbaum et al., 1977). The macerate was transferred to a finisher (fabricated in the Dept. of Food Science & Technology) that was operated
at 1000 pm and with various screen openings (FSO): 0.020, 0.027,
0.033, and 0.045 in. (or 0.508, 0.686, 0.838, and 1.143 mm, respectively). Juice from the finisher was boiled in a steam-jacketed
kettle and hot-filled into No. 303 cans. The cans were sealed, rolled
for 3 min, and spin cooled in cold water. They were stored in a
- 3.9C room until experimentation. Seventy-six cans of tomato juice
from 0.033 in. screen and two cans each of the juice from 0.027 and
0.045 in. screens were produced.

The wet sieving technique proposed by Kimball and Kertesz (1952)


was employed to determined weighted average diameter of particles
in tomato juice and concentrate samples. A set of five U.S.A. Siandard sieve series (Newark Wire Cloth Co., Newark, NJ) with 20, 40,
60, and 100, and 140 mesh openings were used. For the particles
retained on the sieve with the largest openings, the average effective
particle size (diameter) was assumed to be 50% over the diameter of
the openings. For particles which passed through one sieve but not
the next one, an average effective particle diameter half way between
the diameters of the openings of the two sieves was assumed. All
experiments were replicated and the weighted average diameters of
the particles were calculated.

Preparation

Rheological

of tomato concentrates

by evaporation

of juice (JE)

Lots of juice from the finisher equipped with different FSO were
transferred to a steam-jacketed vacuum kettle described in detail by
Saravacos and Moyer (1967). The kettle was operated at 132.4 kPa
(26-27 in. vacuum). The concentrates from 0.033 in screen were
taken out periodically at the approximate total solids (T.S.) of 10, 15,
20, 25, and 28%; in addition, a sample with the highest concentration
of about 30% T.S. was obtained. Other concentrates, from 0.020,
0.027, and 0.045 in. screens, were taken when the concentrations
were about 30%
T.S. The tomato concentrates were canned and
stored as described for the juice.
Preparation
(SE)

of tomato concentrates

by evaporation

of serum

Canned tomato juice from 0.033 in. screen described earlier was
centrifuged at 11,700 X g at 20C for 45 min (Sorvall RC-5, Ivan
Sorvall, Inc., Norwalk, CT), and the volume of serum was measured
and transferred to a beaker. The pulp was scraped from the centrifuge
tubes and transferred into a plastic bottle and stored in a refrigerator.
The serum was concentrated in a steam-jacketed kettle to various
Brix. When the concentrated serum was cooled to room temperature,
it was proportionally combined with the separated pulp to obtain concentrates (150 mL) of 10, 12, 14, 16, 18, 22, and 28 Brix.
A 25 Brix serum concentrate was prepared to study the effect of
heat applied to serum on the rheological properties of the reconstituted
concentrates. It was diluted with distilled water to obtain serum samples at 22, 20, 18, 16, 14, 12, and IO Brix. The diluted serum
samples were proportionally combined with the separated pulp to obtain 100 mL samples of concentrates. All SE concentrates were allowed to rehydrate overnight in the refrigerator. The portions of the
concentrates that were not used on the next day were stored at -3.9C.
Preparation of tomato concentrates
concentration of serum (SRO)

by reverse osmosis

Serum and puip were separated from canned tomato juice in the
manner described for SE concentrates. Instead of evaporation, reverse
osmosis was used to concentrate the serum using a cellulose acetate

membrane(type S-97 CAB) (Osmonics,Inc., Minnetonka,MN) in a


batch type reverse osmosis unit with a volumetric capacity of 200 mL
at 63.1 MPa (900 psig) (Abcor Inc., Cambridge, MA). Continuous
agitation of the serum above the membrane surface was provided by
means of a magnetic stirrer. The concentrated serum was collected
periodically at various Brix and then proportionally combined with
the separated pulp. The SRO concentrates were allowed to rehydrate
in the refrigerator overnight before use in experiments. Portions of
the concentrates that were not used in the experiment on the next day
were stored at - 3.9C.
142-JOURNAL

OF FOOD

SCIENCE-Volume

52, No. 1, 1987

Total solids determination


A sample was weighed in an aluminum pan and dried in a vacuum
oven (Central Scientific Co., Chicago, IL) operated at 2l.lC (70F),
64.9 kPa (28 in. vacuum) for 48 hr. The dry sample was cooled in a
desiccator for at least 2 hr before its weight was measured.
Determination

of particle size distribution

measurements

Flow properties of the concentrates were determined with a concentric cylinder viscometer (Haake RV2, Haake Inc., Saddle Brook,
NJ) as described earlier (Vitali and Rao, 1984) at five temperatures:
lo, 25, 40, 55, and 70C for the concentrates processed from FSO
of 0.045, and 0.033, and 0.027 in. For concentrates from 0.020 in.
screen, whose rheological behavior was determined first, the temperatures employed were 5, 15, 25, 35, and 45C. Yield stress of
samples was determined using the relaxation technique described by
Van Wazer et al. (1963).

RESULTS & DISCUSSION


Rheological properties of tomato concentrates and tomato
juice
Flow curves consisting of log shear rate (y ) against log shear
stress (7) of tomato concentrates from the three different concentration processes and from the four different screen sizes,
as well as concentrated serum, showed power-law behavior.
T = Ky

(1)

Linear regression analysis was performed on the data resulting


in values of slopes (n), intercepts (K), and correlation coefficients. The correlation coefficients were in the range 0.97 to
1.00. The flow behavior index n of the tomato concentrates
was found to vary from 0.266 to 0.444. With values of n being
less than 1, tomato concentrates are shear thinning fluids. The
flow behavior index showed no definite trends with concentration, temperature, or methods of concentration in accordance with the findings by Harper and El Sahrigi (1965) and
Rao et al. (1981). Using magnitudes of K and n, apparent
viscosities of the concentrates were calculated from the relationship:
~)a,100= K (loo)-

(2)

Effect of temperature. The effect of temperature on the


apparent viscosity of the concentrates at 100 set- was described well by the Arrhenius relationship:
WOO

= v= exp WRT)

(3)

Magnitudes of the activation energy (E,) of the concentrates


ranged from 2.0 to 3.0 kcallmole and were within the range
of values reported by others (Harper and El Sahrigi, 1965; Rao
et al., 1981).

whole tomato juice, both serum and pulp are subjected to heat.
Structure of pulp may be affected by heat. Particle sizes or
volume of the pulp may be reduced during the heat treatment.
Moreover, concentrating tomato juice and tomato serum by
heating to the same Brix requires different heating time because tomato juice has lower heat transfer coefficient than the
serum (Kopelman and Mannheim, 1964).
Kopelman and Mannheim (1964) found that SE concentrates
had much lower viscosity than JE concentrates. They concluded that lower consistency in SE may be attributed to the
centrifugation during serum separation (which was not specified in their publication) which led to crushing of the cells and
the disruption of the solid suspension structure of the juice.
However, their tomato concentrates were made by cold break
method (60C). Pectic enzymes may still have been active in
the concentrates resulting in subsequent loss of consistency.
Effect of heat in concentration step. It has been known
for a number of years that when tomato concentratesare diluted
to lower concentration, the diluted products have lower viscosity than if they are concentrated straight from the juice. In
the present study, this effect was first observed for a JE concentrate. Figure 5 contains the apparent viscosities of two 16%
T.S. JE tomato concentrates: one prepared by straight concentration of juice and the other by dilution of a concentrate with
41% total solids. It is clearly seen that the straight concentrate
had higher apparent viscosity than the diluted concentrate.
Figure 6 shows that SE tomato concentrates prepared from
dilution also have lower apparent viscosity than the straight
concentrates. In this case, only the serum experienced heat.
Structure of the pulp should be the same in both concentrates,
diluted and straight, only the nature of the serum was different.
Heat alters the structure of pectic substancesby means of hydrolysis. Colloidal properties of serum may be altered by heat
resulting in lower apparent viscosity of reconstituted tomato
concentrates with unheated pulp. In this respect, Caradec and
Nelson, (1985) reported that viscosity of tomato juice serum
decreasedwith heat treatment. The observation of Caradec and
Nelson (1985) is in agreement with the present results in that
heat treatment reduces the viscosity of serum and juice.
Marsh et al. (1977) found that pulp lost bound water as a
result of the physical forces that developed as concentration
progressed and the loss altered their ability to influence consistency. Therefore, water removal by means of evaporation
may irreversibly affect the rheological properties of the final
products.
Labuza (1977) suggested that the apparent loss of consistency or viscosity was most likely due to the failure of the
macromolecular polymeric substances, comprising the water
insoluble solids, to resorb to their maximum extent. Pectic
substancesand other long-chair carbohydrate polymers can be
hydrolyzed by heat (Kertesz, 1951) resulting in smaller molecules. Colloidal properties exhibited by pectic substancesare
changed. Cell wall materials become less rigid and smaller in
size when heat is applied.

Effect of concentration. The relationships between apparent viscosity and concentration were of the power type. The
exponents did not vary much with either screen sizes or temperatures (Table 1). At 25X, for all screen sizes the exponent
of the power relationship was 2.24 with a correlation coefficient of 0.954; this magnitude is in the range of values: 2.5
and 2.0 reported by Rao et al. (1981) and Harper and El Sabrigi
(1965), respectively.
Effect of screen size
Shear rate-shear stress data of a 20% T.S., JE concentrate
are shown in Fig. 1. From the data, it can be ascertained that
in general smaller FSO (theoretically smaller particle size distribution) yielded lower apparent viscosities. However, the
concentrates from 0.027 in. screen had the highest apparent
viscosity among the four screen sizes. Similar results were
obtained for tomato juice (Somsrivichai, 1986). It is interesting
to note that particle size distributions of samples from 0.027
in. and 0.045 in. screens were similar to each other on one
hand (Fig. 2) and those using 0.020 in and 0.033 in. were
similar to each other on the other hand (Fig. 3).
The observed influence of screen size may be explained in
that small screens reduce the size of the particles. However,
at the same time they remove some of the large particles from
the finished products resulting in tomato concentrateswith narrow particle size distribution and a small amount of large particles. Based on theories of suspensionrheology (Jinescu, 1974),
small suspendedparticles may give high viscosity due to their
greater surface area. Large particles contribute to high viscosity also. Therefore, small screen sizes can affect the gross
viscosity of tomato concentrates in two opposite manners: one
is enhancing gross viscosity due to large surface area of small
particles and the other one is diminishing the gross viscosity
due to the exclusion of large particles. Screen size of 0.020 in
may produce tomato concentrates with too small particle size
distribution and very small amount of large particles resulting
in small magnitude of viscosity while 0.027 in. screen may
produce small particles as well as allow some of the large
particles to be in the tomato concentrates. It may be that using
0.027 in. screen resulted in tomato juice and concentrates with
appropriate particle sizes which yielded the highest viscosity.
Effect of methods of concentration
Effect of methods of concentration on qloo of tomato concentrates with 16% T.S. from three different processes:juiceevaporation, serum-evaporation, and serum-reverse osmosis,
can be seen in Fig. 4. At low concentrations, apparent viscosities of SRO and SE concentrates were not significantly different At higher concentrations, concentrates from serumevaporation were exposed to heat for longer periods of time,
therefore, their apparent viscosities were less than that of concentrates from serum-reverseosmosis. Nevertheless, both SRO
and SE concentrates showed higher apparent viscosities than
that of JE concentrates at the same concentrations.
From data in Fig. 4 it appears that concentrating tomato
serum by means of evaporation or reverse osmosis does not
have significant effect on apparent viscosity of reconstituted
concentrates with unheated pulp. When heat is applied to the
Table l-Slope

Juice
Juice
Juice
Juice
Serum
Serum

Process
evaporation
evaporation
evaporation
evaporation
evaporation
reverse osmosis

of the plot in (q,&

Screen
size
(in.)
0.020
0.027
0.033
0.045
0.033
0.033

Yield stresses of tomato concentrates


Yield stressesof tomato concentrates from juice evaporation
process using four FSO were determined over the concentra-

versus In (total solids) of tomato

concentrates
Temperature

5
2.29
-

10
2.55
2.86
2.77
-

15
2.63
-

25
2.64
2.46
2.94
2.80
2.36
2.82

35
2.82
-

from different

processes

(C)
40
2.50
2.91
2.77
-

45
2.85
-

q
-

55
2.64
3.08
2.84
-

70
2.82
2.97
3.16
-

alq~ is apparent viscosity at a shear rate of 100 x-1.

Volume 52, ho. 1, 1987-JOURNAL

OF FOOD SCIENCE-143

EFFECT OF PROCESSING

ON TOMATO VISCOSITY. .

Tomato Juice Solids


JE
*
5.5

0
A

Concentrates,
20
0.020
in. screen
0.027
0.033

XT.S.,

25

c
E
\
z

0 ScreenSize 0.508mm
+- - High or Low Value
0 ScreenSize 0.838mm
8 - High or Low Value

a 80
E

60

c
3
4.5

t
41

)
4

4.5

5.5
Ln

(f.

6.5

xc-)

100

Fig. l-Shear
rate ($-shear stress (T) data at 25 C of 20% juice
evaporated (JE) concentrates that were made using tomato juices
from different finisher screens.

SiZCeSize
Fig. 3-Volume
of pulp retained on sieves for juice samples
from finisher screens of 0.020 in. (0.508 mm) and 0.033 in. (0.838
mm).

Tomato Juice Solids


0.2

8 high value= 160 ,,

80

0 ScreenSize 1.143mm
High or Low Value

zE
\
m
z

-0.3

60

,a
PC
1 40
*

SE Concentrate,

16.62

XT.S.

JE Concentrate.

15.98

2T.S.

-1.3

20

40

0.0028

100

SiefL Size

tion range 9 to 14%. These concentrations were selected in


order that the most sensitive torque measuring head (50 g-cm)
of the viscometer could be used. Their magnitudes, shown in
Fig. 7, depend on total solids of the concentrates as well as
on FSO. Regression analysis of concentration versus In of yield
stress resulted in quadratic equations as found earlier by (Rao
et al., 1981). The use of larger finisher screens resulted in
concentrates with higher yield stress. Also, as total solids of
the concentrates increased, the magnitude of yield stress increased.
When values of yield stress are determined for a fluid, another flow model containing the yield stress term, such as that
of Herschel-Bulkley (Eq. 4), must be employed to fit the viscometric data.
(4)

144-JOURNAL

OF FOOD SCIENCE-Volume

52, No. 1, 7987

0.0038

(K)

Fig. 4-Apparent
viscosity at 100 set- (q,& as a function of
temperature of 16% concentrates made by evaporation of juice,
evaporation
of serum, and reverse osmosis concentration
of
serum.

-0.2
Q

Diluted

Straight

Concentrate
Concentrate

::

0.0028

0.0029

0.003

0.0031
l/T

In Eq. (4), 7 is the shear stress, i, is the shear rate, r0H is the
yield stress, KH is the consistency index, and nH is the Bow
behavior index.
It should be pointed out that magnitudes of rheoiogical parameters obtained from this analysis in which yield stress was
included will be somewhat different from the analysis based

0.00355

0.0033
l/T

Fig. 2-Volume
of pulp retained on sieves for juice samples
from finisher screens of 0.027 in. (0.686 mm) and 0.045 in. (1.143
mm).

T = TOH + KH 9

0.00305

0.0032

0.0033

0.0034

(K)

Fig. &Apparent
viscosity at 700 set- (q,& as a function of
temperature of straight and diluted, 15% total solids, juice evaporated concentrates, from 0.033 in. screen.

on the simple power law model (Eq. 1). However, apparent


viscosities will be the same in both analyses, only the flow

2.2

2.45

2.7
Ln

2.95

(Concentration.

3.2

2T.S.)

Fig. (T-Plot of In concentration (% total solids) versus In apparent viscosity (Pa.s) at 25 C of serum evaporated straight and
diluted concentrates from 0.033 in. screen.

4
I
3.5

c-

0.020

0
n

0.027
0.033

0.045

in.

screen

\
z

2.5

L
:

1.5

1
9

10

11

12

13

Concentration

14

(5T.S.)

Fig. 7-Yield stress (NlmZ) of juice evaporated (JE) tomato concentrates from different finisher screens as a function of concentration (% total solids).

behavior index, n, and the consistency index, K, will be different.

Dzuy, N.Q. and Boger, D.V. 1983. Yield stress measurement for concentrated suspensions.J. Rheologgy 27(4): 321.
Hand, D.B., Moyer, J.C., Ransford, J.R., Hening, J.C. and Whittenberger,
R.T. 1955. Effect of processingconditions on the viscosity of tomato juices.
Food Technol. 9: 228.
Harper, J.C. and El Sahrigi, A.F. 1965. Viscometric behavior of tomato
concentrates. J. Food Sci. 30: 470.
Herschel, W .H. and Bulkley, R. 1926. Measurement of consistency as applied to rubber-benzene solutions. Proc. Am. Sot. Test. Mater. 26(H): 621.
Jinescu, V.V. 1974. The rheology of suspensions. Int. Chem. Eng. 14(3):
397.
Kattan, A.A., Ogle, W .L., and Kramer, A. 1956. Effect of processvariables
on quality of canned tomato juice. Proc. Am. Sot. Hort. Sci. 68: 470.
Kertesz, Z.I. 1951. The Pectic Substances.Interscience Publishers, Inc.,
New York.
Kertesz, Z.I. and Loconti, J.D. 1944. Factors determinin the consistency
of commercial canned tomato juice. NYSAES Tech. BuK1. 272.
Kimball, L.B. and Kertesz, Z.I. 1952. Practical determination of size distribution of suspended particles in macerated tomato products. Food
Technol. 6: 68.
Kopelman, I.J. and Mannheim, H.C. 1964. Evaluation of two methods of
tomato juice concentration. I. Heat-transfer coefficients. Food Technol.
18: 907.
Labuza, T.P. 1977. The pro erties of water in relationship to water binding
in foods: Review. J. FooB Proc. Preserv. 1: 167.
Luh, B.S., Dempsey, W .H., and Leonard, S. 1954. Consistency of pastes
;a:, puree from Pearson and San Marzano Tomatoes. Food Technol. 8:
Luh, B.S., Leonard, S.J., and Phaff, H.J. 1956. Hydrolysis of pectic materials and oligouronides by tomato polygalacturonase. Food Res. 21: 448.
Mannheim, H.C. and Ko elman, I.J. 1964. Evaluation of two methods of
juice concentration. II. K roduct evaluation. Food Technol. 18: 911.
Marsh, G.L., Buhlert, J., and Leonard, S. 1977. Effect of degree of concentration and of heat treatment on consistency of tomato pastes after dilution. J. Food Proc. Preserv. 1: 340.
McColloch, R.J., Nielson, B.W., and Beavens, E.A. 1950. Factors influencing the uality of tomato paste. II. Pectic changes during processing.
Food Tee% nol. 4: 339.
Mizrahi, S. and Berk, Z. 1972. Flow behavior of concentrated orange juice:
Mathematical treatment. J. Text. Studies 3: 69.
Nelson, P.E. and Tressler, D.K. 1986. Tomato juice and tomato juice blends.
In Fruit and Vegetable Juice Processing Technology, Ch. 12, 3 rd ed.,
AVI Publishing Company, Westport, CT.
Rao, M.A., Bourne, M.C., and Cooley, H.J. 1981. Flow properties of tomato
concentrates. J. Text. Studies 12: 521.
Rao, M.A. and Cooley, H.J. 1983. Applicability of flow models with yield
for tomato concentrates. J. Food Proc. Eng. 6: 159.
Robinson, W .B., Kimball, L.B., Ransford, J.R., Moyer, J.C., and Hand, D.B.
1956. Factors influencmg the degree of settling m tomato juice. Food
Technol. 10: 109.
Saravacos, G.D. and Moyer, J.C. 1967. Heating rates of fruit products in
an agitated kettle. Food Technol. 21: 372.
Smit, C.J.B. and Nortje, B.K. 1958. Observation on the consistency of tomato paste. Food Technol. 12: 356.
Sornsrivichai, T. 1986. A study on rheological properties of tomato concentrates as affected by concentration methods, processing conditions
and pulp content. Ph.D. thesis, Cornell Univ., Ithaca, NY.
Surak, J.G., Matthhews, R.F., Wang. V., Padua, H.A., and Hamilton, R.M.
1979. Particle size distribution of commercial tomato juices, Proc. Fla.
Stats Hort. Sot. 92: 159.
Tanford, C. 1961. Physical Chemistry of Macromolecules.John Wiley,
New York.
Van Wazer, J.R., Lyons, J.W., Kim, K.Y., and Colwell, R.E. 1963. Viscosity and Flow Measurement. A Laboratory Handbook of Rheology.
Interscience Pub., New York.
Vitali, A.A. and Rao, M.A. 1984. Flow roperties of low-pul concentrated
orange juice: Serum viscosity and efpect of pulp content. B Food Sci. 49:
676.

REFERENCES
Birnbaum, D.G., Leonard, S., Heil, J.R., Buhlert, J.E., Wolcott, T.K., and
Ansar, A. 1977. Microbial activity in heated and unheated tomato serum
concentrates. J. Food Proc. Preserv. 1: 103.
Caradee, P.L. and Nelson, P.E. 1985. Effect of temperature on the serum
viscosity of tomato juice. J. Food sci. 50: 1497.
Casson, N. 1959. A flow equation for pigment-oil sus ensions of the printing ink type. In Rheology of Disperse Systems, EC. Mill (Ed.), p. 82.
Pergamon Press, New York.
Charm, S.E. 1962. The nature of role of fluid consistency in food engineermg applications. Adv. Food Res. 11: 356.
Davis, R.B., DeWeese, D., and Gould, W .A. 1954. Consistency measurements of tomato puree. Food Technol. 8: 330.

Whittenberger, R.T. and Nutting, G.C. 1957. Effect of tomato cell structures on consistency of tomato juice. Food Technol. 11: 19.
Whittenberger, R.T. and Nutting, G.C. 1958. High viscosity of cell wall
suspensionsprepared from tomato juice. Food Technol. 12: 420.
Ms. received 4118186;revised 8/22/86; accepted 8/22/86.

Oneof us (TS)was the recipientof a scholarshipfrom the AnandhaMahidolFoundation, Bangkok,Thailand.This work also was supported by funds from the Hatch
act.
Based on a paper presented at the 46th
Technologists,
June 15-18, Dallas, TX.

Volume 52, No. 1, 1987-JOURNAL

Annual

Meeting

of the Institute

of Food

OF FOOD SCIENCE-145

You might also like