You are on page 1of 11

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

net/publication/229706424

Bigelow's General Method Revisited: Development of a New Calculation Technique

Article  in  Journal of Food Science · July 2006


DOI: 10.1111/j.1365-2621.2003.tb09646.x

CITATIONS READS
24 3,442

3 authors:

Ricardo Simpson Sergio Almonacid


Universidad Técnica Federico Santa María Universidad Técnica Federico Santa María
198 PUBLICATIONS   2,628 CITATIONS    75 PUBLICATIONS   1,478 CITATIONS   

SEE PROFILE SEE PROFILE

Arthur A Teixeira
University of Florida
103 PUBLICATIONS   1,705 CITATIONS   

SEE PROFILE

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

Triticale seed properties View project

FONDECYT REGULAR 1140817 View project

All content following this page was uploaded by Ricardo Simpson on 04 September 2018.

The user has requested enhancement of the downloaded file.


JFS: Food Engineering and Physical Properties

Bigelow’s General Method Revisited:


Development of a New Calculation Technique
R. SIMPSON, S. ALMONACID, AND A. TEIXEIRA

ABSTRACT: This article describes the development of a comprehensive procedure for broader application of the
General Method in calculating thermal processes. In addition to the capability of calculating lethality (or process
time to achieve a specified lethality) for a given set of heat-penetration data, the procedure can also calculate
process time or lethality for process conditions different from those of the original heat-penetration tests. The new
procedure was experimentally tested against the Ball Formula Method and numerical finite difference methods. It
often safely predicted shorter process times than the Ball Formula Method by accurately accounting for slow come-
up and cool-down phases of the retort process.
Keywords: general method, formula method, process evaluation techniques

Introduction Bigelow’s procedure earned the name over time (beyond the initial lag) appears as

T HERMAL PROCESSING IS AN IMPORTANT


method of food preservation in the
“General” method because it applies to any
product/process situation. Since it relies
a straight line that can be described mathe-
matically by a simple formula, and related
Food Engineering and Physical Properties

manufacture of shelf-stable canned foods solely on the measured cold spot tempera- to lethality requirements by a set of tables
and has been the cornerstone of the food- ture, it is blind to process conditions, mode that must be used in conjunction with the
processing industry for more than a centu- of heat transfer, product properties, or con- formula.
ry. Thermal process calculations, in which tainer size and shape. This “immunity” to However, there are several assumptions
process times at specified retort tempera- product/process conditions has always made that cause the method to lose accura-
tures are calculated to achieve safe levels of been the strength of the General Method, in cy in many situations. According to Hold-
microbial inactivation (lethality), must be addition to its unquestioned accuracy. For sworth (1997), most Formula Methods have
carried out carefully to ensure public health this same reason, the greatest limitation of been applied to metallic cans or glass jars
safety. However, overprocessing must be the General Method was that it could be that can be processed in pure steam or wa-
avoided because thermal processes also used only to calculate process times for the ter-cook retorts with rapid come-up-times.
have a detrimental effect on the quality (nu- same retort temperature used in the heat Recent developments with retortable flexi-
tritional and sensorial factors) of foods. penetration test from which the cold spot ble pouches and semirigid bowls and trays
Therefore, the accuracy of the methods used temperature profile was obtained. Thus, it have made it necessary to reexamine pro-
for this purpose is of importance to food sci- has limited predictive power (Pham 1987). cess calculation methods. These packages
ence and engineering professionals working Over time, several improvements were in- are processed with steam-air mixtures in the
in this field. troduced to the original General Method, system and often require relatively slow
The first procedure to calculate thermal such as those contributed early on by Ball come-up times, which can introduce addi-
processes was developed by W.D. Bigelow in (1928) and Schultz and Olson (1940), and tional error with use of formula methods.
the early part of the 20th century and is usu- then later by Patashnik (1953) and Hayaka- Most workers in this field will agree that
ally known as the General Method (Bigelow wa (1968). the General Method is more accurate than
and others 1920). The General Method The lack of programmable calculators or the Formula Method, but the popularity of
makes direct use of the time-temperature personal computers until the latter part of the Ball Formula Method as a tradition
history at the coldest point to obtain the le- the 20th century made this method very throughout the food-canning industry con-
thality value of a process. The procedure long, tedious, and impractical for most rou- tinues to be overwhelming (Merson and
was carried out graphically using a plot of tine applications, and it soon gave way to others 1978). According to Teixeira (1992),
lethal rate against time to produce a lethal- formula methods offering shortcuts. In re- the limiting factors that historically deterred
ity curve, the area beneath which corre- sponse to this need, a semi-analytic meth- the use of the Bigelow General Method
sponded to the accumulated lethality deliv- od for thermal process calculation was de- have long since been overcome with the
ered by the process. If more or less lethality veloped and proposed to the scientific advent of programmable calculators and
were required, the procedure was repeated community by Ball (1923). This is the well- personal computers.
with an estimate of the cooling portion of known Formula Method, and works in a dif- The goal of this study was to reintroduce
the cold spot temperature (cooling profile) ferent way from the General Method. It the General Method as a more accurate,
advanced or retarded on a trial-and-error makes use of the fact that the difference powerful, and easy-to-use method of ther-
basis until the desired lethality was between retort and cold spot temperature mal-process calculation. Specific objectives
achieved. This is the reason why this meth- decays exponentially over process time after were as follows:
od was known as the graphical trial-and-er- an initial lag period. Therefore, a semiloga- • Develop a procedure that would integrate
ror method (Stumbo 1973). rithmic plot of this temperature difference the lethality calculation by the General

1324 JOURNAL OF FOOD SCIENCE—Vol. 68, Nr. 4, 2003 © 2003 Institute of Food Technologists
Further reproduction prohibited without permission

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1324 7/23/2003, 12:02 PM
General method revisited . . .

Table 1—Can dimensions and processing conditions for Figure 1 to 7 Table 2—Process lethality and process
time calculated from experimental
Can size Can TRTa,b Pt Fo – Valuec data presented in Figure 8 using RGM
Figure (mm) (common name) (°C) (min) (min) and FM
1 52 × 38 70 g tomato paste 116 50 9.8* Evaluation Fo – Value 1 Pt 2
2 74 × 116 Nr 1 tall 121 90 28.5* Technique (min) (min)
3 99 × 141 Liter 121 95 8.5*
4 52 × 38 70 g tomato paste 116 28 3.1** FM 3.59 111
5 52 × 38 70 g tomato paste 116 13 0.2** RGM 6.31 102
6 99 × 141 Liter 121 70 2.0** 1 F value calculated from experimental data.
o
7 74 × 116 Nr 1 tall 130 40.1 6.0 2 Calculated for a F = 6.3 (min).
r
a TRT(t) 5% Bentonite, 1-kg cans (98 × 110 mm)
= a + bt, ( 0 < t < CUT ), where a = 40 (°C) and b can be evaluated, in each case, considering the j h = 1.93, f h = 71.73 (min)
known value of a and that TRT(5) is the value reported as processing temperature per process. IT = 17.8 °C, TRT =121 °C, T r = 121 °C,
b TRT = Constant ( t > CUT ).
cCalculated with General Method. T w = 24.2 °C
*Fp > Fr
** F p < F r
Fr = 6 min; CUT = 5 min; Conduction heated product , ␣ = 1.7 ⫻ 10 –7 m/s 2 ; T w = 18 (°C), IT = 70 (°C)
rately predicted by integrating the heat-
transfer concepts developed here. Much of
the significance of the work reported here
Method with principles of the heat- The true profile of these cooling curves will stems from the heat-transfer concepts de-
transfer theory. be altered on the basis of when the onset of veloped in the following section.
• Demonstrate its ability to evaluate pro- cooling occurs because different internal
cesses at different conditions from those temperature distributions change continu- Heat-transfer concepts
used in heat-penetration tests (retort ously during heating. At best, these cooling Most mathematical models for predicting
temperature, initial temperature, and so temperature profiles are only crudely and time-temperatures histories in food prod-
on). conservatively estimated in the traditional ucts at a given point normally need to as-
• Demonstrate its ability to take into ac- use of the General Method. Herein lies one sume one of the basic modes of heat trans-
count slow come-up and cool-down

Food Engineering and Physical Properties


of the existing weaknesses that was ad- fer. Two extreme cases have their own
phases. dressed by integrating heat-transfer con- analytical solutions: (1) perfect mixing of a
• Demonstrate that the procedure performs cepts to more accurately predict the true al- liquid (forced convection), and (2) homoge-
with at least the same ease of use and ternate cooling temperature profiles. neous solids (pure conduction). Most foods
reliability as the Formula Method but The third capability has not been possible are an intermediate case, and these ex-
with better accuracy. with traditional use of the General Method. treme solutions would give a guideline for
This application requires accurate predic- the usefulness of temperature-time histo-
Description and identification of tion of the entire cold spot temperature pro- ries (profiles) developed here.
the problem file under totally different conditions of re- Heat-transfer model for perfect mixing.
A detailed procedure was developed to tort and/or initial product temperatures, For forced convection (agitated liquids), it is
give practical use to the General Method including retorts with unusually slow come- possible to assume that temperature inside
proposed by Bigelow and others (1920) so up times. These profiles can also be accu- the can is uniformly distributed but time-
that it would include the following capabil-
ities: (1) calculation of lethality (Fo value)
for a given set of heat penetration data, (2)
calculation of process time to achieve a
specified lethality (Fr value) from a given
set of heat-penetration data, and (3) calcu-
lation of either process time or lethality for
alternative process conditions different
from those used during the original heat-
penetration tests (even if there is slow
come-up time) with no further experimental
data.
The first capability is a straightforward
execution of the General Method with no
involvement of heat transfer. The accumu-
lated lethality is calculated by numerical
integration of the lethal rate along the cold
spot temperature profile measured in the
heat penetration test (or provided by simu-
lation of a test).
The second capability requires execution
of the first as a starting point. Integration
must then be repeated with the cooling por-
tion of the cold spot temperature profile
advanced or retarded on a trial-and-error Figure 1—Simulated heat-penetration data for analysis (Fp > F r) for Case 1—
basis until the desired lethality is achieved. Situation 1

JFS is available in searchable form at www.ift.org Vol. 68, Nr. 4, 2003—JOURNAL OF FOOD SCIENCE 1325

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1325 7/23/2003, 12:02 PM
General method revisited . . .

Table 3—Process lethality and process must be the same at different TRT and/or
time calculated from experimental IT:
data of a broken-heating curve pre-
sented in Figure 9 using RGM and FM (7)
Evaluation Fo – Value1 Pt2 (4)
technique (min) (min)
FM 35 26 From Eq. 7, the dimensionless tempera-
RGM 40 25 ture ratio can also be expressed as:
Slow come-up time with perfect mixing.
1 F value calculated from experimental data.
o
2 Calculated for a F = 6.0 (min). Eq. 5 was derived from Eq. 2, solving an or-
r
Product: meat chunks, UT can 73 × 115 (mm) dinary differential equation and assuming
j h = 5.65, f h = 12.85 (min), f h2 = 47.96 (min)
IT = 46.5 °C, TRT =127.9 °C, T r = 121 °C, a linear retort temperature profile (that is,
T w = 25.9 °C simulating temperature profile during
come-up time).

dependent. A transient energy balance, tak-


ing the container as a system, gives: (8)

(1) (5)
Heat transfer model for pure conduction.
Heat transfer for pure conduction is based
where retort temperature is time-depen- on Fourier’s equation and can be written as:
(2) dent and expressed as: TRT (t) = a + bt and
Eq. 5 is valid for: 0 < t ⱕ CUT. For t > CUT,
Provided that the can’s inside tempera- temperature T (or TC.P.) can be expressed by (9)
ture is uniformly distributed, T also de- Eq. 3 using an appropriate initial tempera-
Food Engineering and Physical Properties

notes the cold spot temperature (T = TC.P.). ture (constant TRT). If thermal conductivity (k) is indepen-
Using the initial condition as T = IT at t = 0, Provided that fh is defined as ln10 ⫻ [MCp/ dent of temperature and the food material
and T at time t > 0, the integration of Eq. 2 UA] (Merson and others 1978), Eq. 5 can be is assumed isotropic, as it is for most foods
renders: rearranged and expressed as: at the sterilization temperature range, then
Eq. 1 becomes:

(3)
(10)
The dimensionless temperature ratio for (6)
forced convection (Eq. 3) is dependent on Although solutions for different geome-
geometry, thermal properties, and time. Further working on Eq. 6 renders: tries are not necessarily straightforward, in
Therefore, the liquid’s aforementioned ratio general, for any geometry, the dimension-
less temperature ratio for constant retort
temperature can be expressed as (Carslaw
and Jaeger 1959):

(11)

If initial temperature distribution, geom-


etry, product (thermal properties), and time
are maintained constant (just changing TRT
and/or IT), then the dimensionless tem-
perature ratio of the solid must be the same
at different TRT and/or IT:

(4)

It is important to point out that Eq. 11 is


valid for constant retort temperature (TRT);
so is Eq. 4. A simplified analytical solution
for homogeneous solids confined in a finite
cylinder is presented in Eq. 12 (Merson and
Figure 2—Simulated heat-penetration data for analysis (Fp > F r) for Case 1—
Situation 2 others 1978). This simplified solution is only

1326 JOURNAL OF FOOD SCIENCE—Vol. 68, Nr. 4, 2003 JFS is available in searchable form at www.ift.org

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1326 7/23/2003, 12:02 PM
General method revisited . . .

valid for long periods of time (after the ini- Table 4—Can dimensions and processing conditions for experimental data
tial lag period when the Fourier number is shown in Figure 10
greater than 0.6), in addition to assuming a TRT F o1 (min) F o1 (min) Pt2 (min) Pt2 (min)
Biot number over 40 (meaning that the ex- Process (°C) IT (°C) RGM FM RGM FM
ternal heat resistance is negligible in com- 1 114.5 17.3 7.0 6.0 36 40
parison with the internal resistance). 2 117.6 63.1 13.6 9.9 16 23
1 F value calculated from experimental data.
o
2Calculated for a F = 6 (min)
r
Product: mussel ( Mytilus chilensis ) in brine, can format 100 ⫻ 22 (mm)
f h = 9.11 (min) a; j h = 1.21a
a Obtained from the process at TRT = 114.5 (°C)

(12)

Slow come-up time with conduction


heating. Gillespy (1953) and Hayakawa
(1974) have developed methods to deter-
mine center temperature where the heating
profile was time-dependent (for example,
linear or exponential). According to Hold-
sworth (1997), the method is applicable to
packs, which have a slow come-up, for ex-
ample, conduction heating products in flex-

Food Engineering and Physical Properties


ible pouches or plastic containers. Gillespy
(1953) developed an equation for a slab of
material being heated with a linear temper-
ature gradient valid during come-up time.
Hayakawa (1974) developed a similar
expression for finite cylinders. Expressions
for conduction heating products of other
geometries (for example, parallelepiped)
with a linear temperature gradient can be
found in Carslaw and Jaeger (1959) and Lu-
Figure 3—Simulated heat-penetration data for analysis (Fp > F r) for Case 1—
ikov (1968). Situation 3
According to Carslaw and Jaeger (1959)
and Luikov (1968) it is possible to find a di-
mensionless temperature ratio equation
suitable for a linear heating profile during
come-up time in conductive heating prod-
ucts.
Heat-transfer model: a general ap-
proach. Although the heat-transfer mecha-
nisms are rather dissimilar, both models
(pure conduction and forced convection),
within certain limitations, can be described
by the same mathematical expression that
was presented by Ball (1923):

(13)

Where:

As was shown by Datta (1990), the latter


expression is valid not only for finite cylin-
ders, but also for arbitrary shapes (rectangu- Figure 4—Simulated heat-penetration data for analysis (Fp < F r) for Case 2—
Situation 1

JFS is available in searchable form at www.ift.org Vol. 68, Nr. 4, 2003—JOURNAL OF FOOD SCIENCE 1327

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1327 7/23/2003, 12:02 PM
General method revisited . . .

lar, oval shape, and so on). The main limita- ature ratio for conduction-heating products. may be bigger, smaller, or equal to Fr. There-
tions are that for heat conduction foods, it is These statements have been supported by fore, 3 situations arise: Fp > Fr, Fp = Fr, and
only valid for heating times beyond the ini- computer-aided experiments that demon- Fp < Fr. Situations such as Fp > Fr and Fp < Fr
tial lag period (when the Fourier number is strate that Eq. 8 is an accurate and secure are of interest for further analysis and they
greater than 0.6). way to transform data obtained for conduc- will be called Case 1 (Fp > Fr) and Case 2
An interesting, practical, and general con- tion-heating products subjected to a linear (Fp < Fr ).
clusion that can be drawn from the heat- temperature profile during come-up time. Case 1 are all those heat-penetration
transfer theory presented here is that Eq. 4 The importance and relevance is that we tests in which final lethality is bigger than
remains independent of container geome- will be able to transform the raw data from the required lethality, so the processing time
try and heat-transfer mode (conduction or heat penetration tests and use the General must be shortened, for example, to find a
forced convection) and requires only a con- Method, not only to directly evaluate the new processing time shorter than the real
stant retort temperature: raw data, but also to evaluate processes at processing time, so that Fp ⱖ Fr and Fp – Fr is
different conditions (retort temperatures, a minimum. Within this case, 3 different sit-
initial temperatures, longer or shorter pro- uations may arise. Figures 1, 2, and 3 were
(4)
cess times) than those originally recorded. computer-generated to show and analyze
the 3 different situations (specifications giv-
and for the cooling phase: Methods and Materials en in Table 1).
In the first situation (Figure 1), shorten-
Thermal process evaluation (to ing the process time was straightforward
(14) calculate lethality) because with the new process time (for an F
The method must allow for the calcula- value equal to or bigger than Fr), the tem-
tion of the F value for a set of data obtained perature inside the can is uniform and it can
Although Eq. 4 is valid only for constant
experimentally or by simulation. Given that be assumed that the cooling temperature
retort temperature profiles, Eq. 8 has shown
the data are not continuous, the integration profile would be the same as the original. In
that similar expressions for the dimension-
procedure should be done numerically the second and third situations, the prob-
less temperature ratio can be derived for
(Gauss, Simpson, trapezoidal, and so on) or lem is different and more complicated.
Food Engineering and Physical Properties

the case of slow come-up time (for example,


alternatively fit the data by interpolation In Figure 2, the temperature at the cold-
linear temperature rise during come-up time
method (such as cubic spline) and integrate est point (for the adjusted process) would
for forced convection heating products).
the lethality analytically. be lower than retort temperature, giving a
Even though Eq. 8 was derived for situa-
tions in which forced convection is the dom- non-uniform temperature distribution in-
Thermal time adjustment (to side the can. In this situation, the tempera-
inant heating method—so as to use a single
calculate process time) ture at the cold spot (referred to as the heat-
equation for data transformation—this one
To determine the processing time so ing part) during the cooling phase will have
will also be used on products in which the
that the F value obtained (Fp) is equal or inertia. To evaluate process lethality (for the
ruling heating mechanism is conduction. As
greater than the required lethality (Fr), the adjusted process), it is necessary to generate
was previously mentioned by some authors
Fp value had to first be determined, with the data for the cooling phase. In this case, the
(Carslaw and Jaeger 1959; Luikov 1968), it is
original heat penetration data. This Fp value use of Eq. 14 will generate data assuming no
feasible to derive a dimensionless temper-
inertia. Although this is not completely ac-
curate, the resulting error in predicted le-
thality coming from the cooling phase will
be on the conservative (safe) side.
In Figure 3, similar to Figure 2, the tem-
perature at the cold spot for the process is
lower than the retort temperature as well as
for the adjusted process. Although, both
curves have inertia, again the use of Eq. 14
will not be accurate but will safely predict
lethality from the cooling phase for the ad-
justed process. In this application, the use of
Eq. 14 will predict inertia but less pro-
nounced than in the real curve.
Case 2 includes all those heat-penetra-
tion tests in which final lethality is lower
than required, so the processing time must
be extended, that is, find a new processing
time longer than the test processing time, so
that Fp ⱖ Fr and Fp – Fr is a minimum. In this
case, the cooling phase temperature profile
must be displaced to the right in order to
extend process time. Three situations need
to be considered and analyzed; they are
Figure 5—Simulated heat-penetration data for analysis (Fp < F r) for Case 2— presented in Figure 4, 5, and 6 (specifica-
Situation 2 tions given in Table 1). In the first situation

1328 JOURNAL OF FOOD SCIENCE—Vol. 68, Nr. 4, 2003 JFS is available in searchable form at www.ift.org

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1328 7/23/2003, 12:02 PM
General method revisited . . .

(Figure 4), the process adjustment is case of Figure 7, two aspects should be care- than the original). In this case, the trans-
straightforward. The two assumptions fully considered when changing processing formed data would overestimate the F value
needed are that the coldest temperature conditions (TRT’ and/or IT’): (1) maintain that could result in an unsafe process. In
could be maintained at the same level and the same come-up time and (2) decide how this situation (which is very rare; Figure 7
that the new cooling temperature profile to generate the new cooling temperature presents an extreme situation), a safe proce-
would be the same as the original one. profile, specifically in the presence of differ- dure should be to follow the recommenda-
In Figure 5 and 6, it was first necessary to ent inertia. First, although it is a limitation, tion explained in Case 2, Figure 5 and 6.
generate more data for the heating process come-up time must be maintained. Sec- To manage the already transformed data
to be able to extend the process time using ond, the cool-down temperature data trans- (new processing conditions), it is necessary
Eq. 13, followed by the need to generate the formation could lead to a new cooling tem- to follow the procedure explained in the sec-
new cooling temperature profile. perature profile with less inertia (for tion “Thermal time adjustment (to calculate
According to heat-transfer theory, the example, the new retort temperature is lower process time)” to adjust the process.
recorded cooling temperature profile will
have more pronounced inertia than the new
cooling temperature profile; therefore, in
these 2 situations (Figure 5 and 6), the use
of Eq. 14 would generate not only inaccurate
but also unsafe data. To avoid this problem,
Eq. 14 was applied, considering only the
data starting at point A as seen in Figure 5
and 6 (A pinpoints the maximum tempera-
ture recorded in the process). Since these
data lack inertia, they could generate the
new cooling temperature profile on the safe
side.

Food Engineering and Physical Properties


Thermal-process evaluation at
conditions other than those
experimentally recorded (or
generated by simulation)
Sometimes it is useful to obtain a pro-
cess evaluation at different conditions other
than those used for the original heat-pene-
tration test and avoid or significantly re-
duce the number of new experiments. The
new process conditions could be: initial food
temperature, retort temperature, and/or Figure 6—Simulated heat-penetration data for analysis (Fp < F r) for Case 2—
cooling temperature. The new time-temper- Situation 3
ature data should be predicted using ade-
quate mathematical models (if the type of
food allows it) or using the dimensionless
temperature ratio concept developed in this
study that is applicable to any kind of food.
The dimensionless temperature ratio con-
cept could be used for any kind of geometry.
However, in real process situations, the re-
tort temperature is not always constant (for
example, come-up time) and will impair the
theoretical validity of the concept derived
for dimensionless temperature ratio, as has
been discussed in the literature (Shultz and
Olson 1940). In the present work the retort
temperature was divided into 3 parts: (1)
come-up time (TRT(t) = a + bt), (2) process
temperature (TRT = Constant), and (3) cool-
ing temperature (Tw = Constant).
Eq. 8 was used to transform the original
data to the newest processing conditions
(TRT’ and/or IT’), assuming a linear temper-
ature profile during come-up time; and Eq.
4 and 14 were used for the constant retort
temperature (TRT) and cooling water tem- Figure 7—Simulated heat-penetration data for analysis (changing processing
perature (Tw) conditions, respectively. In the conditions, TRT)

JFS is available in searchable form at www.ift.org Vol. 68, Nr. 4, 2003—JOURNAL OF FOOD SCIENCE 1329

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1329 7/23/2003, 12:02 PM
General method revisited . . .

Table 5—Comparison of operator process time (Pt) required for the same le- Validation
thality at alternative retort temperatures calculated by RGM and FM with ac-
Thermal process evaluation and adjust-
tual operator process time in data set generated by finite differences (F.D.)
(conduction heated product) at each retort temperature. ment. To compare results from the Revisited
General Method (RGM) developed in this
TRT (°C) F.D. (min) RGM (min) Error (%) FM (min) Error (%)
study with those from the Formula Method
110 436 461 5.73 465 6.65 (FM), sets of computer-simulated data as
115 341 367 7.62 370 8.50 well as experimental data were analyzed.
125 245 251 2.45 275 12.24
130 217 227 4.61 247 13.82 Experimental data. Figure 8 shows exper-
imental data taken from Teixeira and others
F r = 6 (min); Conduction heated product, can (603 ⫻ 909); ␣ = 1.25 × 10 –7 (m 2 /s); CUT = 30 (min);
TRT = 120 (°C), IT = 50 (°C), Tw = 20 (°C); f h = 289.2 (min); j ch = 1.8 (1999) for a thermal process that was evalu-
ated with the RGM as well as with the tradi-
tional FM. Figure 9 shows experimental data
for a broken-heating curve (Tucker 2002).
Results of process evaluations (Figure 8 and
9) by both procedures as well as adjusted
processes for a specified F value are depict-
ed in Table 2 and 3.
Simulated data. To analyze extreme situ-
ations, heat penetration data (cold spot
temperature profiles) were generated at dif-
ferent retort temperatures using a finite dif-
ference solution of the conduction heat-
transfer equation for a cylindrical can
(Teixeira and others 1969) and for forced
convection product using Eq. 3 and 5 that
Food Engineering and Physical Properties

were developed in this study. The data set


generated at 120 °C was then used as a start-
ing point (reference process) to calculate
process times to achieve a specified lethal-
ity by both methods. Calculations with the
FM were executed with computer software
publicly available at the Purdue Univ. Food
Science Dept. website <http://cifmc.
foodsci.purdue.edu/ball/ball.cfm>. Calcu-
lations with the RGM (thermal process eval-
Figure 8—Experimental data for a heat-penetration test (taken from Teixeira
and others (1999) uation at different conditions and adjust-
ments) were executed according to the
procedure described previously in the
Methodology section.
Thermal process evaluation at different
conditions other than Recorded and Ad-
justment for a specified Fo value. Figure 10
represents experimental data collected in a
seafood-processing plant to test the devel-
oped procedure for changing retort temper-
ature and/or initial temperature and then
adjusting the process to a specified Fr value.
Process specifications and calculations are
given in Table 4.

Results and Discussion

Thermal-process evaluation and


adjustment
Experimental data. First, for the situation
shown in Figure 8, application of the RGM is
no different from the original General
Method because no data transformation
was involved, and the data were directly
evaluated. Table 2 shows the results from
evaluating experimental data using the
Figure 9—Experimental data for a heat-penetration test (broken heating curve). RGM and the FM. The FM severely under-
(Tucker 2002) estimated the Fo value whereas the RGM

1330 JOURNAL OF FOOD SCIENCE—Vol. 68, Nr. 4, 2003 JFS is available in searchable form at www.ift.org

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1330 7/23/2003, 12:03 PM
General method revisited . . .

produced an estimate in very close agree- Table 6—Comparison of operator process time (Pt) required for the same le-
ment. In this example, the FM overestimat- thality at alternative retort temperatures calculated by RGM and FM with ac-
tual operator process time in data set generated by analytical solution (forced
ed operator process time (Pt) by nearly 9% convection product) at each retort temperature.
compared with the RGM. This is a reflection
of the inherent weakness in the FM for eval- Analytical
TRT (°C) solution (min) RGM (min) Error (%) FM (min) Error (%)
uation of lethality at the onset of the cool-
ing phase (Merson and others 1978; Spinak 110 130.9 130.6 –0.23 127.1 –2.90
115 75.9 75.8 –0.13 72.5 –4.48
and Wiley 1982). This is particularly signifi-
125 40.1 40.1 0.00 37.7 –5.99
cant in the situation shown in Figure 8 be- 130 32.0 32.0 0.00 30.1 –5.94
cause the accumulated lethality during the Convection heated product, can dimensions 0.1 m dia ⫻ 0.1 m height
cooling phase is greater than that accumu- Fr = 6 (min); U = 100 (W/m 2 °C); TRT = 120 (°C), IT = 50 (°C), T w = 20 (°C); CUT = 30 (min); fh = 45.3
(min); j ch = 1.2
lated during the heating phase.
Figure 9 represents experimental data
for a broken heating curve. Again the FM
underestimated the accumulated lethality
compared with the RGM (Fo = 35 min com-
pared with 40 min). When adjusting process
time for an Fo value of 6 min, the RGM gave
a slightly shorter operator process time
(Pt = 25 min compared with 26 min).
Simulated data. Computer-supported
experiments were developed to analyze ex-
treme situations. Extreme cases were se-
lected as (1) being pure conduction and
forced convection, (2) having extreme fh val-
ues, and (3) having a slow come-up time.

Food Engineering and Physical Properties


Conduction-heated product. Table 5
compares process times for 4 alternative
processes (reference process was developed
at 120 °C). Product (thermal diffusivity) and
can dimensions were chosen to have a high
fh value (289.2 min). In addition, the refer-
ence process considered a very high come-
up time (30 min). In all 4 computer-simu-
lated experiments, the RGM had less error
than the FM. Note that the error in the
RGM, in this example, was insensitive to re- Figure 10—Experimental data obtained in seafood-processing plant for a heat-
penetration test at 2 retort temperatures
tort temperature. On the other hand, in the
case of the FM, the error was significantly
higher as retort temperatures approached
130 °C, as others have shown (Holdsworth
1997; Smith and Tung 1982). Figures 11 and
12 show graphically how the RGM was ap-
plied to generate and adjust the new pro-
cess starting with the reference process at
TRT = 120 °C. Figure 11 depicted results
from changing retort temperature from 120
to 130 °C. First, the come-up cold spot tem-
perature profile was transformed using Eq.
8. Second, for the period of constant retort
temperature, Eq. 4 was used. Third, given
that the reference process (TRT = 120 °C)
had less inertia at the cooling phase com-
pared with the new process at 130 °C, the
dimensionless temperature ratio concept
was directly applied to the entire cooling
phase. Finally, the new process
(TRT = 130 °C) was adjusted for a Fo value of
6 min according to the procedure described
in methodology. Figure 12 depicted results
from changing retort temperature from 120
to 110 °C. In this case, the data at 120 °C (ref- Figure 11—Thermal process at 130 °C for a conduction-heated product ob-
erence process) were insufficient to achieve tained from a reference process at 120 °C

JFS is available in searchable form at www.ift.org Vol. 68, Nr. 4, 2003—JOURNAL OF FOOD SCIENCE 1331

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1331 7/23/2003, 12:03 PM
General method revisited . . .

a Fo value of 6 min at 110 °C. Then, using Eq. processes (reference process was developed ature. It is also important to note that in
13, more heating data were generated at at 120 °C). Given that Eq. 6 was derived for these 4 examples, the FM predictions re-
120 °C before proceeding with data transfor- perfect mixing (linear temperature profile sulted in an error found on the risk side.
mation to 110 °C. To transform thermal data during come-up time), the dimensionless
from 120 to 110 °C, Eq. 8 was used for come- temperature ratio (Eq. 8) could accurately Thermal process evaluation at
up, Eq. 4 for holding time, and finally Eq. 14 transform the original time-temperature conditions other than those
for the cooling phase but starting from point data (for 0 < t < CUT). Equation 4 was used recorded, and adjustment for a
A (Figure 12), avoiding inertia to safely pre- for t > CUT, and Eq. 14 for the cooling-down specified Fo value
dict the cooling phase at 110 °C. phase. As seen in Table 6, the error attribut- Experimental data presented in Figure
Convection-heated product. Table 6 ed to the RGM was approximately zero in all 10 were selected as being normal thermal
compares process times for 4 alternative 4 cases, regardless of the new retort temper- processing data. First, come-up between
both processes was similar but with slight
differences. Second, initial temperatures
(IT) were different; and third, the retort
temperature (TRT = 117.6 °C) was constant
but with slight variations during the process.
Thermal process data were evaluated with
the RGM and the FM (Fo values are depicted
in Table 4). In both cases, the FM underesti-
mated Fo value in relation to the RGM, re-
vealing once again that the higher the retort
temperature, the lower the prediction ca-
pacity of the FM.
Taking Process 1 (Table 4) as a reference
process, thermal process data were trans-
formed from TRT = 114.5 °C to
Food Engineering and Physical Properties

TRT = 117.6 °C according to the RGM proce-


dure and then used for process evaluation to
compare with results from the FM. When
comparing operator process time (Fr = 6
min), FM overestimated both processes by
approximately 10% and 44%, respectively,
compared with RGM (TRT = 114.5 °C and
TRT = 117.6 °C). According to Figure 13, the
RGM has a very accurate prediction capacity
and the predicted Fo value is on the safe
Figure 12—Thermal process at 110 °C for a conduction-heated product ob- side (Fo value of 6.07 min compared with
tained from a reference process at 120 °C
6.14 min).

Conclusions

I T WAS POSSIBLE TO DEVELOP A SYSTEMATIZED


procedure for the General Method to give
it the same ease of use as the Ball’s Formula
Method. The RGM allows the same calcula-
tions as the FM, but with more accuracy.
The proposed methodology was rather
accurate when used for thermal process
adjustment. The processing time was al-
ways estimated (overestimated) with an er-
ror less than 5% (in all cases under study).
For the FM, it was common to find errors of
10% to 20% or more. The new procedure
safely predicted shorter process times than
those predicted by the FM. These shorter
process times may have an important im-
pact on product quality, energy consump-
tion, plant production capacity, and ade-
quate corrections for online control (process
deviations). On the other hand, the FM pre-
dicted unsafe processes for forced convec-
tion products analyzed in this study.
Figure 13—Use of dimensionless temperature ratio concept over experimen- Further testing with experimental data
tal data to generate data at a different retort temperature must be done on this developed procedure

1332 JOURNAL OF FOOD SCIENCE—Vol. 68, Nr. 4, 2003 JFS is available in searchable form at www.ift.org

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1332 7/23/2003, 12:03 PM
General method revisited . . .

to check its performance and then it can be = rate of heat transfer sterilizing value of a thermal process. Food Tech-
nol 22(7):93-5.
made available in computer user–friendly R = inside radius of can Hayakawa KI. 1974. Response charts for estimating
software. The developed procedure could be TA = extrapolated initial can temperature temperatures in cylindrical cans of solid food sub-
jected to time variable processing temperatures. J
extended to pasteurization processes, UHT/ obtained by linearizing entire heating Food Sci 39(6):1090-8.
HTST processes, and so on. Future studies curve of a can Holdsworth SD. 1997. Thermal processing of pack-
aged foods. London: Blackie Academic & Profes-
should consider the possibility of changing T = temperature sional. p 146-61.
come-up time, the nonlinear temperature TC.P. = temperature in the coldest point Luikov AV. 1968. Analytical heat diffusion theory. New
York: Academic Press, Inc. p 300-50.
profile during come up time, and the capa- T’C.P. = new temperature in the coldest point Merson RL, Singh RP, Carroad PA. 1978. An evalua-
bility to evaluate broken heating curves at dif- IT = initial temperature tion of Ball’s formula method of thermal process
ferent conditions than the originally record- ITW = initial temperature cooling phase calculations. Food Technol 32(3):66-76
Patashnik M. 1953. A simplified procedure for ther-
ed temperature (TRT and/or IT). IT’ = new initial temperature mal process evaluation. Food Technol 7(1):1-6
IT W⬘ = new initial temperature cooling Pham QT. 1987. Calculation of thermal process le-
thality for conduction-heated canned foods. J Food
Nomenclature phase Sci 52(4):967-74.
A = area TRT = retort temperature Schultz OT, Olson FC. 1940. Thermal processing of
canned foods in tin containers. III. Recent improve-
a and b = constant of linear equation de- TRT ‘: new retort temperature ments in the General Method of thermal process
scribing retort temperature profile Tw = cooling temperature calculation. A special coordinate paper and meth-
ods of converting initial retort temperature. Food
[TRT(t) = a + bt] Tw’ = new cooling temperature Res 5:399.
b⬘ = new slope of the linear equation de- Tr = reference temperature, 121.1 (°C) Smith T, Tung MA. 1982. Comparison of Formula
Methods for calculating thermal process lethality.
scribing retort temperature profile t = time J Food Sci 47(2):626-30.
[TRT’(t) = a’ + b’t] U = global heat transfer coefficient Spinak SH, Wiley RC. 1982. Comparisons of the Gen-
a⬘ = new constant of the linear equation de- eral and Ball Formula Methods for retort pouch
process calculations. J Food Sci 47(3):880-4, 888.
scribing retort temperature profile Greek letters Stumbo CR. 1973. Thermobacteriology in food pro-
[TRT’(t) = a’ + b’t] ␣: = thermal diffusivity of food (␣ = k/␳Cp) cessing. 2nd ed. New York: Academic Press. p 143-
51.
Cp = heat capacity of food ␳: = density of food Teixeira AA. 1992. Thermal process calculations. In:
CUT = come-up time ⵜ = differential or nabla operator Heldman DR, Lund DB, editors. Handbook of food
engineering. New York: Marcel Dekker, Inc. p 563-
= energy per mass unit (ⵜ = )

Food Engineering and Physical Properties


619.
Fo = sterilizing value at 121.1 °C ⵜ2 = laplace operator Teixeira AA, Balaban MO, Germer SPM, Sadahira MS,
Teixeira-Neto RO, Vitali AA. 1999. Heat transfer
Fp = process sterilizing value (ⵜ = ) model performance in simulation of process devi-
Fr = required sterilizing value ations. J Food Sci 64(3):488-93.
f = rate factor (related to slope of semi-log References Teixeira A, Dixon J, Zahradnik J, Zinsmeiter G. 1969.
Computer optimization of nutrient retention in the
heat-penetration curve) Ball CO. 1923. Thermal processing time for canned
thermal processing of conduction-heated foods.
foods. Bull. Nr 7-1 (37). Washington, D.C.: Natl. Res.
fh and fc = heating and cooling rate factors Food Technol 23(6):845-50.
Council.
Tucker GS. 2002. Personal Communication. Chipping
(related to slope of semi-log heat-pene- Ball CO. 1928. Mathematical solution of problems on
Campden, Gloucestershire, U.K.: Campden and
thermal processing of canned food. Berkley, Calif.:
tration curve) Chorleywood Food Research Assn.
Univ. Cal. Pub. In Pub. Health 1, N 2, 15-245.
MS 20020308 Submitted 5/20/02, Revised 12/8/02,
Bigelow WD, Bohart GS, Richardson AC, Ball CO. 1920.
Accepted 1/22/03, Received 1/27/03
j = dimensionless lag factor Heat penetration in processing canned foods. Bull.
Nr 16-L Res. Washington, D.C.: Lab. Natl. Canners We kindly appreciate the contribution made by Mr. Cristian
jh and jc = heating and cooling lag factors Assn. Cortés and particularly to Mrs. Paula Solari.
Carslaw HS, Jaeger JC. 1959. Conduction of heat in
k = thermal conductivity of food solids. London: Oxford Univ. Press. p 63-4. Authors Simpson and Almonacid are with the
l = height of canned content Datta AK. 1990. On the theoretical basis of the as- Dept. de Procesos Químicos, Biotecnológicos, y
ymptotic semi logarithmic heat penetration curves Ambientales, Univ. Técnica Federico Santa María;
M = product mass used in food processing. J Food Eng 12:177-90. P.O. Box 110-V; Valparaíso, Chile. Author Teixeira
Pt = operator process time (time that is mea- Gillespy TG. 1953. Estimation of sterilizing values is with the Dept. of Agricultural and Biological
of processes as applied to canned foods. II. Packs Engineering, Frazier Rogers Hall, P. O. Box
sured from when the retort reaches pro- heating by conduction: complex processing condi- 110570, Univ. of Florida, Gainesville, FL 32611-
cessing temperature [TRT] until the tions and value of coming-up time of retort. J Sci
0570. Direct inquiries to author Simpson (E-mail:
steam is turned off ). Food Agric 4:553-65.
Hayakawa KI. 1968. A procedure for calculating the
ricardo.simpson@pqui.utfsm.cl).

JFS is available in searchable form at www.ift.org Vol. 68, Nr. 4, 2003—JOURNAL OF FOOD SCIENCE 1333

jfsv68n4p1324-1333ms20020308-TLS-3pgsColor.p65
1333 7/23/2003, 12:03 PM

View publication stats

You might also like