Anda di halaman 1dari 22

c 

    

 
     
  
  
 
 
 
       
  
    
   
   


    

The following article has been condensed and adapted from Lipid Oxidation by Edwin N.
Frankel, Chapter 12, "Frying Fats." For more information, see this article's end.--Eds.

Deep-fat frying is perhaps the most complex of food processing operations because of the
multiplicity of reactions that occur and the vast quantity of chemical products that are generated.
Understanding the chemical processes that occur during frying and the methods available to
control oil degradation can have a big impact on finished product flavor and shelflife.

Frying is usually done in the 180[degrees]-200[degrees]C range. There are three major chemical
reactions that occur: oxidation, polymerization and hydrolysis. Each of these reactions produces
a wide range of compounds which contribute to the formation of both desirable and undesirable
products in fried foods.

Thermal lipid oxidation and hydrolysis produce complex mixtures of volatile and non-volatile
monomeric and polymeric substances. Polar materials are polymers, which are useful indicators
of oil quality. The food itself also releases proteins, carbohydrates, fats, oils and metals into the
frying medium that may promote reactions affecting oil performance as well as its health and
nutritional properties.

The rapid decomposition of hydroperoxides and aldehydes during frying produces an intricate
assortment of volatile compounds. However, they are found in very low concentrations because
the frying process acts like a large steam distillation operation. Those that remain are important
because many at extremely low levels (in parts per million) contribute to the distinctive flavors
and aromas so typical of fried foods.

The chart "Various Volatiles" shows some of the many compounds found in frying oils and/or
French fries. One of the compounds listed in this table, 2,4-decadienal, a decomposition product
of linoleate that produces rich, fried food flavor, was a significant flavor noted during sensory
evaluation of French fries. The presence of tins linoleate in vegetable oils contributes to the
formation of 2,4-decadienal. When frying potato chips, research indicates that a certain level of
linoleate is essential for desirable flavor development. The downside is that the presence of
linoleate also increases the probability that rancidity will occur. Because decadienal imparts a
desirable "fried flavor" at low concentrations, but too much can cause an undesirable "rancid"
flavor, the frying process also must be carefully controlled by a turnover process to introduce
fresh make-up oil during frying.

 


Both fats and oils of animal and vegetable origin are used for frying. General considerations
involved in the selection of frying oil include the turnover rate (i.e., oil added to a fryer to
replace that removed by the fried products), the amount of fat absorbed by the food and the
shelflife of the food. For high-volume frying, oils should not exceed 2%-3% linolenic acid, a
polyunsaturated fatty acid that is readily oxidized and polymerized. The process of partial
hydrogenation was aimed at reducing the linolenic acid content of soybean and canola oils.
However, partial hydrogenation has the problem of producing trans isomers that are nutritionally
undesirable. Many frying fats are developed by blending oils with much lower degrees of
hydrogenation to meet the recent FDA requirement of labeling the content of trans fats in foods.
In response to this issue, plant breeders have produced zero-trans oils with higher oleic acid
content and reduced linolenic acid to improve stability during frying without hydrogenation.

There are many ways to measure the changes that occur in oils during frying. As frying
progresses, the oil darkens, becomes more viscous and tends to foam, and the degree of
unsaturation drops. Chemical methods which estimate the amount of polar materials, total
polymeric materials and/or free fatty acids also are routinely used to monitor oil degradation.

Polar materials are measured by column chromatography. This is a time-consuming, expensive


method and consequently is not suitable for quality control. To reduce costs and minimize the
amount of solvent used, rapid micro chromatographic methods have been developed.

The rate of polar material formation varies with the type of frying operation. Oils used for
restaurant frying can contain up to 30% polar materials after 40 hours of frying. The reason for
this rapid buildup is that batch frying is done intermittently and oil turnover is slow (i.e., oil is
not frequently replaced). Turnover in industrial frying operations is much greater due to the
volume of food being prepared. The frying fat generally reaches a steady-state condition after a
few hours of continuous operation. The amount of polar material produced by the three common
frying operations can be compared in the chart "Frying Operations and Amount of Polar
Compounds."

There also are a number of rapid test methods available. These include a range of commercial
colorimetric tests based on redox indicators, carbonyl compounds, free fatty acids and alkaline
materials. Unfortunately, these methods are not specific, and [this author suggests they are not to
be] recommended as tools for quality monitoring. The ultimate arbiter of frying, however, is in
the taste and quality of the fried food. The most sensitive and reliable methods for monitoring
frying oil degradation are based on sensory evaluation of the foods and gas chromatographic
analysis of the volatile flavor compounds in the foods, preferably potatoes fresh and after
storage.

c  

A number of control mechanisms are available to enhance food quality and extend the usable
frying life of the cooking oil. These include monitoring and controlling frying oil temperatures,
increasing turnover rates, filtering the oil, using additives such as antioxidants and
polymerization inhibitors and controlling fat uptake by the foods.

Optimal frying is conducted at temperatures ranging from 160[degrees]-190[degrees]C, and


frying at lower temperatures will minimize thermal degradation of the oil. However, sufficiently
hot oil is essential to ensure that food is thoroughly cooked. The continuous emanation of
bubbles from frying food is proof that moisture is escaping as steam. Steam release helps strip
volatile decomposition products from the oil, which delays fat deterioration. To reduce the
potential for thermal degradation and oil breakdown, the temperature of the oil can be lowered
when not in use.

Filtration is used to remove charred food particles and batter that accumulate in the oil. Removal
of these materials can slow oil degradation, maintain oil color and minimize the development of
off-flavors. Metal screens, paper and plastic are used as filter materials. Filtration can be
augmented by using filter aids such as diatomaceous earth.

The use of additives can enhance both oil life and the shelflife of flied foods. Antioxidants such
as BHA, BHT, propyl gallate and TBHQ may be included in commercially used frying fats and
oils. These oils also may include naturally occurring or added tocopherols, which generally are
depleted during the frying process. Natural antioxidants such as rosemary and sage also may be
added to flying oils. Like the synthetic antioxidants noted above, these also can protect the oil
during frying and enhance shelflife of fried food.

The amount of fat absorbed by the food during frying affects its sensory perceptions and
acceptability. The chemical makeup of the frying oil and that picked up by the food are not
significantly different, and minimizing the rate of oil degradation can reduce the amount of oil
picked up by the food.

Deep-fat frying is perhaps the most complex processing operation known. The frying oil is a
dynamic entity that changes depending upon the food being fried and the time of frying. And
although oil degradation helps produce some of the delicious flavors that make fried foods so
palatable, understanding its processes and the means to control them is essential to producing
healthy and nutritious food.

This article was condensed and adapted, in part, by Rick Stier from Lipid Oxidation, 2nd ed.,
written by Edwin N. Frankel, University of California, USA, and published by The Oily Press,
PJ Barnes & Associates. Published March 200.5. 486 pages, 152 tables, 148 figures, 87
equations, 849 references. Volume 18 in The Oily Press Lipid Library. For more information,
call: 44-1823-698973; e-mail: sales@pjbames.co.uk. See also www.pjbarnes.co.uk/ol)/lo2.htm.

Various Volatiles

Compounds Found Description (b)


in French Fries (a)

Hexanal Green
trans, trans-2,4-decadienal Deep-fried
2-octenal Tallowy, nutty
Nonanal Fatty, green
trans-2-nonenal Tallowy, green
cis, trans-2,4-decadienal Deep-fried
Dimethyltrisulphide Cabbage
2-ethyl-3,6-dimethylpyrazine Roasty, earthy

Source: Derived from Table 12.2, pg. 359, Lipid Oxidation,


The Oily Press, Bridgwater, England

a) From Belitz and Grosch, 1986

b) From Wagner and Grosh, 1997


COPYRIGHT 2006 BNP Media
COPYRIGHT 2008 Gale, Cengage Learning

http://users.auth.gr/~karapant/tdk/Research/Project15c.html

Ò   


     
  


Contact person:  !
" [lioumbas@gmail.com] and  #
 [agzamanis@yahoo.gr]

Transport through porous media has gained great attention in the last decades due to its relevance to a wide range of
applications such as electronics cooling, thermal insulation engineering, water movement in geothermal reservoirs,
heat pipes, underground spreading of chemical waste, nuclear waste repository, geothermal engineering, grain
storage and enhanced recovery of petroleum reservoirs. In many of the aforementioned problems of interest, the
porous medium consists of single-phase (O  ) as well as two-phase (O    
) regions as a result of
evaporation due to vapour pressure i.e. below the liquid¶s boiling temperature. One common and easy practice in
everyday life process which is closely related to transport phenomena through porous media is frying. Frying is an
extremely complex process and involves unsteady state coupled heat and mass transfer problem in porous media
(      
 ), phase change of water, bubble formation and bubbly flow combined with
unsteady state heat transfer in the oil bulk.
During frying we are dealing with ³unconventional´ boiling on the surface of a fluid-saturated porous material, since
the bubbles are not formed solely because of the rapid vaporization of a liquid -heated to its boiling point- on the
porous material surface, but are formed mainly due to the heat transfer from the hot oil (that surround the porous
medium) to the -entrapped within the porous structure- water (  
 
 O   O   ) in a region
close to porous interface ( ). Specifically, the penetrating liquid not only wets and fills the porous network
but also heats it up, while gases or vapors need to escape from the porous network to allow room for liquid
penetration. Such gas/vapor departure occurs in the form of bubbles which depart from the outside surface of the
cold porous material to the surrounding hot liquid. The bubble formation and departure from the porous material as
well as natural convection of the unevenly heated liquid is expected to significantly be affected by gravity. The
problem becomes even more complicated if one takes into account the dramatic changes in the properties of the food
being fried ( 
O 

        
            
   

 
O      
   
      ) and the frying medium (         O
  O  O

  O).  presents schematically the scientific areas interrelated with the
process of frying.
We expect that the experimental results of the present activity (       O  O   
 



      
O  
  O 
 
O O   
 O   
          


     
  O  
  
  
  


O ‘ 
    

  ) in conjunction with modeling efforts would contribute to elucidate the observed phenomena during the
conventional heat and mass transfer in porous media in one hand, and on the other hand will offer a new insight in
studying the conventional frying process with a novel perspective that includes heat and mass transfer in porous
media.


  $ %      
     
    
 
  

     ' 
&     
    

   







International Journal of Signal System Control and Engineering Application Year: 2009 |
Volume: 2 | Issue: 2 | Page No.: 35-39 DOI: 10.3923/ijssceapp.2009.35.39

( ( 


   )
   


Driss Izbaim, Bouaazza Faiz, Ali Moudden, Naima Taifi and Adil Hamine

http://www.medwelljournals.com/fulltext/?doi=ijssceapp.2009.35.39

"  
The main objective of this study is to evaluate quality changes of Soybean Oil (SBO) during
frying. Traditional methods are expensive and time consuming techniques, requiring technical
personnel and laboratory facilities. Fatty Acids (FA) value, total Polar Component (PC) and Free
Fatty Acids (FFA) of oil samples were used as chemical indicators of different quality levels of
oil and then compared with ultrasonic measurements. The study demonstrated that using
ultrasonic properties, we obtain reliable results to monitor and control oil quality.

Ò (cÒ

Frying has become the most popular food preparation technology due to its flavour and the easy
preparation. Important characteristics of industrial frying oils are high oxidative stability, high
smoke-point, low foaming, low melting point, bland flavour and nutritional value. However,
complex physical and chemical changes occur during deep-fat frying leading to thermal and
oxidative decomposition. Physical changes are mainly increased viscosity and foaming, colour
changes and decreased smoke-point. Main chemical changes are increased free fatty acids and
polar components as well as decreased levels of unsaturation, flavour quality and nutritive value
(Warner, 2002).

Pokorny (1989) has demonstrated that increases in the polar fraction resulted in an eventual
degradation in food quality. Many countries consider the polar compounds measurement to be
the single most important test for the degradation state of oil and have established the value of
around 25% as a its regulatory limits in frying oils (Firestone, 1996), some other countries also
use free fatty acids. Free fatty acids produced in the frying operation of an oil contribute to the
smoke haze and therefore has a substantial effect on its smoke point, which affects oil absorption
by the fried food (Orthoefer O., 1996).

There are several methods for controlling and assessing degradation of oil during deep frying,
but are time-consuming, costly and usually require analytical expertise. In this respect, ultrasonic
techniques are fast, non-invasive and convenient. Many studies have been conducted to assess
the composition of different types of food products using ultrasound (McClements, 1997; Mulet
O., 1999). Velocity is the ultrasonic parameter most used because of its reliable results.

The objective of this study is about the changes in ultrasonic measurements during
thermoxidation of SBO and the relationships between ultrasonic measurements and chemical
results.

Ò! 

SBO was obtained from Cristal S.A. (Casablanca, Morocco). The oils was heated in a domestic
fryer at 180°C for 8 h day-1 over 4 days, for a total of 32 h. Samples of 150 mL were periodically
removed and kept at -18°C for further analysis.

( 
   The experimental setup (Fig. 1) used for the experiments consisted
of ultrasonic transducer (5 MHz, 0.5 inch crystal diameter, A309S-SU model, panametrics,
Olympus), attached to a cubic container (50x50x50 mm), where the oil samples were placed. The
container was introduced into a temperature controlled bath to maintain the sample temperature
and the oil was moderately stirred to avoid formation of bubbles. The ultrasonic measurements
were carried out, while the oil sample was cooled from 50-30°C. The transducer was linked to a
pulser-receiver (Sofranel Model 5073 PR, Sofranel Instruments), which sent the electrical signal
to a digital storage oscilloscope (LeCroy 9310 M, LeCroy Cor).

The ultrasonic impulse is propagated in water and crosses the oil sample contained in the vessel
before reflecting on a Plexiglas surface of plate 2 (Fig. 2). Figure 3 shows, the experimental
signal composed of echoes A1 to A4.
Fig. 1: Experimental set-up

Fig. 2: Schematic of echoes reflected by the sample


(A1 to A6 are observed echoes reflected
from the interfaces between media Mi)

Fig. 3: Typical waveform of the reflected ultrasonic signal


with the 5 MHz transducer
For ultrasonic velocity measurement, eight signal acquisitions were taken and averaged. It can be
written as:

(1)

with Ȧ = 2ʌv

To obtain the phase experimentally, the FFT of signals A2 and A4 are calculated (Bakkali O.,
2001).

The attenuation coefficient Į was computed by fitting the experimental data to equation:

(2)

Where:

d = The distance travelled by the wave

A0 = The initial amplitude of the signal measured as the peak-to-peak voltage

A = The amplitude of the signal at distance (d)

Eight ultrasonic echoes were considered to compute attenuation.

c 
 

  
 The FA profile analysis was performed by derivatization to their corresponding
methyl esters (Hartman and Lago, 1973) prior to the analysis by GC. Oil samples (50 mL) are
diluted in the hexane to obtain a solution to be analyzed (about approximately 0.1%).

A VARIAN 3800 chromatograph on a CP Select CB (VARIAN) capillary column (50 m x 0.25


mm i.d., film thickness 0.25 ȝm), was used under the following temperature program: 185°C (40
min), 15-250°C/10 min. Samples injection is 1 ȝL (split ratio 1:100) at 250°C and the flow rate
of Helium, used as carrier gas, was 1.2 mL min-1. Temperature of both split injector and flame
ionization detector was 250°C.

   
 FFA content as the percentage of oleic acid was determined using AFNOR NF
T 60-204 standard method. Acid value was defined as the amount (mg) of KOH required to
neutralize FFA in 1 g of oil sample dissolved in a mixture of diethyl ether and ethanol in the
presence of phenolphthalein.
   The content of total polar compounds was determined following the method
proposed by the IUPAC (1992).

(! Òc(Ò

c 
    

 Table 1 shows the results for the Fatty Acids (FA) profiles of the
SBO carried out on the oil samples at different heating periods. The FA profile of the frying oils
changed as a result of cyclization, polymerization and hydrolytic, oxidative and other chemical
reactions promoted by frying conditions (Nawar, 1996). The degradation affected unsaturated
fatty acids more than saturated fatty acids. The FA composition of particular oil has marked
effects on its frying performance as well as on its physical and chemical behaviour. The linoleic
acid level in deep-frying oils appears to be an obviously negative factor in oil stability. Indeed,
previous studies indicated that a lowered linoleic acid content in soybean oil resulted in
improved oil quality during cooking and frying (Tompkins and Perkins, 2000). Changes in the
FA profile during frying provide only limited information about these compositional changes,
which are associated with oil degradation.

The regular limits in many countries for PC on frying fats and oils are around 25% and in some
others are around 0.3% for FFA as the official regulations.

During oxidation and hydrolysis, FFA are formed, as a result of the cleavage of triglycerides
(Perkins and Erickson, 1996). FFA content is the most frequently used test, but it is not
recommended to use it as the only indicator of oil quality. Oils with high FFA are known to have
a lower smoke point (Augustin O., 1987) and the surfactant effect of FFA contributes to the
foaming, which leads to further oxidation of the oil. Previous studies of frying oils have shown
that the FFA content increases during deep-frying (Kalapathy and Proctor, 2000). As expected,
FFA content of the SBO increased significantly during frying and after 20 h, its content is 0.31%
(Table 2).

Many researchers consider measurement of TCP to be one of the most reliable indicators of the
state of the oil deterioration (Fritch, 1981; Gere, 1982). According to Billek  O. (1978) and
Paradis and Nawar (1981), polar compounds indicate the degradation of oils and the breakdown
of triglycerides. Table 2 shows that TPC in SBO increased significantly during frying. After 32 h
of frying, the final TPC level is 25%. The TPC-based stability is 32 h of frying for SBO not to
exceed the established limit. This would have occurred at even shorter time if the oil was used
for frying foods swamp.

( 
    The ultrasonic velocity and attenuation depend on the physico-
chemical properties of the medium. Ultrasonic velocity decreases with the temperature in fat
(McClements, 1997). Figure 4 shows the influence of temperature on the ultrasonic velocity
measurements, for 8, 16, 24 and 32 h.

Table 1: Soybean oil FA compounds at several


frying times (%)
Table 2: Soybean oil polar compounds and free
fatty acids at several frying times (%)

Fig. 4: Variation of velocity with temperature for


different frying times

Fig. 5: Variation of velocity with time of cooling


for different frying times
As expected, the ultrasonic velocity decreases in line as the temperature increases. The average
velocity temperature coefficient is -3.78 m/sec/°C. On the other hand, the ultrasonic velocity
increases as the time of cooling increases (Fig. 5).

The time of cooling allows making the difference between the various frying. Every sample of
the oil is differently cooled from others according to the duration of frying. Therefore, we can
distinguish two fryings without using the temperature.

Fig. 6: Relationship between the temperature and


the time of cooling for different frying
times (±1°C)

Fig. 7: Variation of attenuation with temperature for


different frying times
The ultrasonic velocity is related to the temperature, which in its turn is related to the time of
cooling. Figure 6 shows the relationship between the temperature and the time of cooling. This
curve is almost the same one for the various fryings.

The ultrasonic attenuation is also affected by temperature (Fig. 7). However, velocity is the
ultrasonic parameter most used because of its reliable results (Benedito O., 2002).

It is not easy to determine the useful life of oil because, it depends on many factors, especially
the composition of the oil and the type of oil used. Significant polynomial fits were found when
relating the ultrasonic measurements and the chemical parameters.

Figure 8a and b show the possibility to detect the limits of the degrees of the polar components
and free fatty acids (PC = 22%, FFA (as oleic acid) = 0.37%) using the ultrasonic measurements.
The ultrasonic measurements are related to the chemical results, which show the feasibility of
using an automated ultrasonic system to monitor the oil quality.

Fig. 8a: Relationship between the ultrasonic


measurements and the percentage of free
fatty acids at 30°C
Fig. 8b: Relationship between the ultrasonic
measurements and the percentage of polar
compounds at 30°C

c c!(Ò

The chemical changes are linked to changes in ultrasonic parameters, which show that ultrasound
can be used to assess and monitor oil quality. The velocity evolution obtained according the time
of cooling appears to be also an indicator of oil degradation, in addition to the evolution of
velocity according to the temperature.

The feasibility of using ultrasonic techniques to rapidly evaluate quality parameters for heated
edible oils was investigated.

 
 

)
c  

 

JAOCS, Journal of the American Oil Chemists' Society, Feb 2009 by Aladedunye, Felix A,
Przybylski, Roman
http://findarticles.com/p/articles/mi_hb5762/is_200902/ai_n32316879/

"  

The changes in regular canola oil as affected by frying temperature were studied. French fries
were fried intermittently in canola oil that was heated for 7 h daily over seven consecutive days.
Thermo-oxidative alterations of the oil heated at 185 ± 5 or 215 ± 5 °C were measured by total
polar components (TPC), anisidine value (AV), color components formation, and changes in
fatty acid composition and tocopherols. Results showed that TPC, AV, color and trans fatty acid
content increased significantly (P
Keywords Canola oil * Frying performance * Total polar component * Anisidine value * Color *
Frying temperature * Tocopherols * French fries * Fatty acids

Introduction

Deep-fat frying is probably one of the most dynamic processes in all of food processing.
Essentially, the process involves immersing a food item in a large quantity of heated oil or fat,
which is normally replenished and reused several times before being disposed. Deep-fat frying
produces a product with desired sensory characteristics, including fried food flavor, golden
brown color, and a crisp texture [1].

Most frying operations are conducted at temperatures of 175-195 °C, nevertheless German
regulations allow maximal frying temperatures of up to 165 °C, to limit formation of acrylamides
[2]. Extruded products and pellets are typically fried at 190-215 °C [2]. This high temperature
requirement and the presence of air and moisture, from the food, initiate several chemical and
physical changes affecting oxidative degradation of oil used. Published studies described
chemical reactions involved and various volatile and non-volatile oxidation products were
identified [3-6]. The chemical changes in the frying fats also affect the physical characteristics of
the oil and fried product [7]. For instance, the color of frying oil was reported to darken as a
result of oxidation and the formation of browning pigments when potato chips were fried [8, 9].

A number of studies have been undertaken to assess various chemical reactions and extent of
oxidative deterioration as affected by frying temperature, many of the published data were
obtained by heating an oil but not during actual frying [10-12]. Meanwhile, it has been observed
that the chemical reactions that take place during deep-fat frying are different from those during
continuous heating [13, 14], Besides, different oils have been found to behave differently
regarding the rate of formation of polar components and secondary oxidation products. Guillen
and Cabo [15] reported that secondary products were formed immediately after hydroperoxide
formation in olive and rapeseed oils, whereas in sunflower and safflower oils, secondary
products were formed when the concentration of hydroperoxides reached level of 180 and 270
meq/kg, respectively. Consequently, the need to study the frying performance of individual oil as
a function of frying temperature during actual frying of food becomes imperative.

Oxidized short-chain fatty acids are secondary oxidation products formed through thermal
degradation of lipid hydroperoxides. Recently, much concern has been on the biological effects
of oxidized lipids, and there is increasing evidence that they may be detrimental to health,
especially in connection with the development of atherosclerosis, liver damage, and promotion of
intestinal tumors [16].

The objective of this study was to evaluate the effect of frying temperature on the degradation of
canola oil by monitoring the accumulation of total polar components, oxidized short-chain fatty
acids, polymers formation, p-anisidine value, color components formation, changes in fatty acid
composition, and tocopherol contents.


 
 

 


Oil and French Fries

Commercially refined regular canola oil without antioxidants added was obtained from
Richardson Oilseed Processing (Lethbridge, Canada). Frozen par-fried French fries in an
institutional pack were obtained from a local food store.

c 


All solvents and chemicals of analytical grade used in this study were purchased from Sigma-
Aldrich (St Louis, MO). Standards of tocopherols were obtained from CalbiochemNovabiochem
(San Diego, CA). Standards of fatty acid methyl esters were purchased from Nu-Check-Prep
(Elysian, MN).

 
  
 


The frying was simultaneously conducted in two 8-L capacity restaurant style stainless steel deep
fryers (General Electric Company, New York, USA). Regular canola oil (3.75 L) was heated at
185 ± 5 and 215 ± 5 °C for 7 h daily for 7 days. A batch of 200 g of frozen French fries was fried
for 5 min for a total of eight batches per frying day. At the end of each frying day, fryers were
shut off and left to cool overnight. Two 25-mL samples of oil from each of the fryers were taken
daily and kept frozen at -16 °C until analyzed. Before commencing frying each day, oils were
filtered to remove solid debris. Oil was replenished every second day of frying with 500 mL of
fresh oil.

 



Fatty acids were methylated following the AOCS Official Method Ce 1-62 [17]. The resulting
fatty acid methyl esters (FAME) were analyzed on Trace GC Ultra gas Chromatograph (Thermo
Electron Corporation, Rodano, Italy) using a Trace TR-FAME fused silica capillary column (100
m × 0.25 mm × 0.25 µm; ThermoFisher Scientific, Waltham, MA, USA). Hydrogen was used as
the carrier gas with a flow rate of 1.5 mL min^sup -1^. The column temperature was
programmed from 70 to 160 °C at 25 °C min^sup -1^ and held for 30 min, and further
programmed to 210 °C at 3 °C min^sup -1^. Starting and final temperatures were held for 5 and
30 min, respectively. Splitless injection was made using a PTV injector. Detector temperature
was set at 250 °C. FAME samples, 1 µL, were injected with an AS 3000 autosampler (Thermo
Electron Corporation, Rodano, Italy). Fatty acids were identified by comparison of retention time
with authentic standards. Oxidized short-chain fatty acids methyl esters (OFAME) were
identified as a group as described by Velasco et al. [18]. Trans isomers of fatty acids were
assessed according to ISO method 15304.

  c 




*
TPC were determined by gravimetric method after column chromatography separation of non-
polar fraction following AOAC Method 982.27 [19]. Polar components were eluted from the
column with diisopropyl ether and further analyzed for polar components composition by size
exclusion chromatography.

Anisidine values (AV), a measure of secondary oxidation products, was determined according to
ISO Method 6885:2004 [20].

  

Tocopherols were analyzed by AOCS Official Method Ce 8-89 [17]. Briefly, oil samples (75
mg) were weighed directly into vials and dissolved in 1.5 mL hexane. Analysis was performed
on a Finnigan Surveyor liquid Chromatograph (LC) (Thermo Electron Corporation, Rodano,
Italy) with a Finnigan Surveyor Autosampler Plus and Finnigan Surveyor FL Plus fluorescence
detector, set for excitation at 292 nm and emission 394 nm. The column was a normal-phase
Microsorb 100 silica column (3 µm; 250 × 4 mm; Varian, CA). Of each sample, 10 µL was
injected. Mobile phase consisted of 7% methyl-teri-butyl-ether in hexane with a flow rate of 0.6
mL/min. The amounts of tocopherols were quantified using calibration curves for each isomer
separately.


+,
c    

The composition of polar components was analyzed using high performance size exclusion
chromatography (HPSEC) according to ISO Method 16931 [21]. Separation was performed on a
Finnigan Surveyor LC. Components were separated on three size exclusion columns in series
(Phenogel 500 Å, 100 Å and 50 Å, 5 µm, 300 × 4.60 mm; Phenomenex, Torrance, CA), with
tetrahydrofuran (THF) as the mobile phase at a flow rate of 0.3 mL/min, and column temperature
of 30 °C. A 10 µL sample was injected, and components were detected with a Sedex 75
evaporative light scattering detector (Sedere, Alfortville, France), operated at 30 °C with an air
pressure of 2.5 bar. Polar components were identified according to the method described by
Marquez-Ruiz et al. [22].

c 


Color of the frying oils was determined according to AOCS Official method Cc 13c-50 [17]
using a DU®-65 spectrophotometer (Beckman, Fullerton, CA).

 




Samples from three repetitions of frying at each temperature were collected and were analyzed in
triplicate. Data are presented as mean values ± SD. Data were analyzed by single factor analysis
of variance (ANOVA) and regression analyses using Minitab 2000 statistical software (Minitab
Inc., PA, ver. 13.2). Statistically significant differences between means were determined by
Duncan's multiple range tests. Statistically significant differences were determined at the P

 


The fresh oil had 0.06% of free fatty acids (FFA), 1.0 meq/kg of peroxide value (PV), 4.2% of
polar components and an AV at the level of 4.2, indicating good quality oil [23]. Thus, changes
in these values during frying would indicate a degradation in oil quality.

  c 

The determination of TPC in frying oil provides the most reliable measure of the extent of
oxidative degradation [14, 24]. In this study, the contents of TPC increased almost linearly with
the frying time at a rate affected by frying temperature (Fig. 1). Total polar content during frying
at 185 °C was 19.8% at the end of frying time, which was still below the 24% oil discard level
set in many European countries [25]. However, the total amount of polar components reached the
discard level after 4 days of frying at 215 °C. The TPC reached 38% by the end of the frying
time. The extent of oxidative deterioration, as measured by TPC formation, was 2.6 times faster
during frying at 215 °C compared to 185 °C.

c 

  c  

The composition of polar compounds formed during frying was analyzed using HPSEC.
Diacylglycerides (DG), oxidized triacylglycerides (OTG), dimers and polymers were separated,
and their contribution calculated using peak area. The contribution of polymers in total polar
material increased consistently with frying time at both frying temperatures achieving maximum
values of 8 and 15.6% for frying at 185 and 215 °C, respectively (Figs. 2 and 3). The amount of
polymers generated at 185 °C at the end of the 7 days frying period was comparable to the third
day of frying at 215 °C. Comparable increase in the amount of dimers for oil fried at 185 °C was
observed throughout the frying period (Fig. 2). However, when frying at 215 °C, a 16-fold
increase in the contribution of dimers at the end of first day followed by slight increase for the
next 2 days of frying, then decreased until the end of the frying period (Fig. 3). This is probably
due to the conversion of the dimers to polymers and thermal degradation of these components
[16]. As expected, the contribution of OTG decreased consistently over the frying period at both
tested temperatures, as a consequence of thermal degradation. However, a more pronounced
decrease in the contribution of OTG, 1.5 times faster degradation was observed during frying at
215 °C (Fig. 3).




*

Aldehydes formed during oxidative degradation are secondary decomposition products, and the
non-volatile portion of carbonyls remains in the frying oil [4, 13]. At the two testing
temperatures, AV was not well correlated with frying time (Fig. 4). The maximum was reached
on the second day of frying for both frying temperatures and then decreased consistently until the
end of frying time. Apparently, oil replenishment played some role in the changes in carbonyls
content but elevated temperatures was probably the main cause for the reduction in the amount
of these labile and reactive components with time [26]. This result could be explained by the
thermal degradation of the aldehydes formed at higher temperature, which results in a lower
accumulation in the oil at the higher frying temperature. Regardless of the general knowledge
that the decomposition of hydroperoxides increases with increasing temperature and potentially
the amount of carbonyls, AV followed an opposite trend in the frying tests. Carbonyl chemical
reactivity, involvement in the formation of other compounds and thermal decomposition explains
decreasing in AV [10, 25]. Houhoula et al. [27] reported a significant increase in AV as a
function of temperature during frying potato chips in cottonseed oil. Thus, the initial increase in
AV observed (Fig. 4) agrees with these authors. Furthermore, the increase was also observed as a
function of temperature.

 
c 



The fatty acid composition of the fresh canola oil and the resulted changes during the 7 days of
frying at 185 and 215 °C are presented in Table 1. The results indicate a progressive decrease in
both linoleic and linolenic acids contributions throughout the frying period [24], Linoleic acid
decreased by 8.5 and 13.3% during frying at 185 and 215 °C, respectively. The deterioration of
linolenic acid was more pronounced and was reduced by 24.0 and 47.1% during frying at 185
and 215 °C, respectively. White et al. [24] reported decreases of 7-11.5% in linoleic acid and 27-
46% in linolenic acid when soybean oils were heated at 180 °C for 40 h.

The amount of trans fatty acids formed during frying increased when temperature and time
increased (Figs. 5 and 6). At the lower frying temperature applied, the amount of trans isomers
increased from 2.4 to 3.3%. The increase in frying temperature to 215 °C caused extensive trans
isomerization of fatty acids (Figs. 5 and 6). The total contribution of trans isomers in oil
increased 2.5-fold, from 2.4 to 5.9% (Fig. 5). This indicates the importance of temperature on
trans isomers formation during frying, and explains the amount of trans isomers observed in the
initial oil (Fig. 5). The deodorization of canola oil is usually performed at temperatures above
200 °C under vacuum, where the main amount of trans isomers is formed [28]. The amount of
individual trans fatty acids decreased in the following order: linolenic > linoleic > oleic (Figs. 5,
6). The quantity of trans isomers formed at elevated temperature indicates that a specific amount
of energy is required to transfer double bonds from cis to trans configuration. Data from this
work are supported by published results, confirming that activation energy for isomerization
decreasing when the numbers of cis double bonds increases [29].

Increasing amounts of trans isomers during frying at higher temperature, can have practical
implications related to nutritional claims about zero trans content in a serving portion of fried
products. When the amount of these isomers increases 2.5-fold during frying at the higher
temperature, then the amount of trans isomers in fried products will increase by the same amount
and may exceed the specified definition limit, making the claim for zero trans fat invalid. The
amount of trans isomers in oil can affect the trans level in fried product due to fast exchange of
fats during frying. We observed that the oil in fried food had a similar composition of fatty acids
as frying oil, even when par-frying was done in different oil (data not included). These data
clearly indicate the importance of controlling frying temperature and keeping it below 190 °C.

The ratio of linoleic acid to palmitic acid (C ^sub 18.2^/C^sub 16:0^) has been suggested as a
valid indicator of the level of PUFA deterioration [34]. Our result showed a decrease in this ratio
from 4.73 to 3.87 and 4.28 to 3.23 during frying at 185 and 215 °C, respectively (Table 1). This
implies that the decrease in this ratio was 1.2 times greater in oil heated at 215 °C as compared to
185 °C. Örnal and Ergin [30] reported a decrease in the ratio from 4.04 to 3.49 at the end of
frying time. Houhoula et al. [27] reported a reduction of the ratio from 2.39 to 2.03 for
cottonseed oil heated at 185 °C for 12 h. The decrease in the ratio of linolenic acid to palmitic
acid was more pronounced, reducing it 1.9 times faster in oil heated at 215 °C compared to 185
°C (Table 1).

Short-chain glycerol-bound aldehydes, acids, ketones and alcohols are non-volatile secondary
oxidation products formed during oxidative degradation of lipids [18]. They are of particular
chemical and nutritional interest since they remain in the frying oil, and are absorbed and
subsequently ingested. Analysis of the oxidized short-chain fatty acid methyl esters (OFAME) as
a group revealed a consistent increase in their contribution for the first 5 days of frying at 185
°C, reaching a maximum at 1.83% (Fig. 7). However, for the oil heated at 215 °C, the amount of
oxidized fatty acids increased in the first 3 days of frying with the maximum at 2.20%.
Thereafter, a decrease for the next 3 days of frying was observed. Velasco et al. [31] observed
that the amount of polar FAME decreased drastically as compared to the amount of total polar
compounds in used frying fats.

Peers and Swoboda [32] suggested the quantification of methyl octanoate, product of linoleic
acid degradation, as an oxidation index during frying. In this study, the amount of octanoic acid
increased significantly (P Color Analysis

  

The tocopherol profile of the fresh canola oil used in this study was found to be: 214 µg/g of a-
tocopherol, and 347 µg/g of y-tocopherol. Tocopherols degradation increased as a function of
frying temperature. At the end of the seventh day of the frying period, approximately 31% of the
total tocopherols remained after frying at 185 °C (Fig. 10). For frying at 215 °C the entire
amount of tocopherols was gone at the end of sixth day of frying. The calculated half life of
tocopherols for oil heated at 185 °C was 8 h while for frying at 215 °C was 5.3 h. Consistent
with previously published results [35], a strong inverse relationship was observed between TPC
formation and the reduction of tocopherol at both frying temperatures. In this study, y-tocopherol
degraded at a faster rate than a-tocopherol at the lower frying temperature, but the order was
reversed during frying at 215 °C (data not shown).

c 
- &    

Although TPC remains the best assessment parameter for evaluating frying oil performance and
oxidative stability, a faster and yet objective alternative is desirable. Assessment methods that
correlate well with TPC may provide the much needed alternative. In this study, a low
correlation was found between AV and TPC, and AV and color at both frying temperatures
(Table 2). However, a good correlation was observed between color and TPC at both frying
temperatures (Table 2). Lopez-Varela et al. [35] reported a correlation coefficient of 0.885
between color and TPC for sunflower oil used in 75 successive fryings of potatoes. In summary,
the effect of temperature and frying time on performance of canola oil as measured by TPC, AV,
fatty acid composition, tocopherols amount and color was significant.
c


The rate of the thermal and oxidative degradation of PUFA was drastically higher at the elevated
temperature tested, forming larger amounts of components with potential detrimental health
effects. This study showed that increasing frying temperature above 195 °C can cause intensive
isomerization of PUFA and the amount of trans isomers may increase above the threshold level
described by "zero trans definition" annulling the trans fat free claim for fried product.

  

1. Warner K (2004) Chemical and physical reactions in oil during frying. In: Gupta MK, Warner
K, White PJ (eds) Frying technology and practice. AOCS, Champaign, pp 16-28

2. Gupta MJ (2004) The effect of oil processing on frying oil stability. In: Gupta MK, Warner K,
White PJ (eds) Frying technology and practice. AOCS, Champaign, pp 76-90

3. Chang TY, Eiserich JP, Shibamoto T (1993) Volatile compounds identified in headspace
samples of peanut oil heated under temperature ranging from 50 to 200°C. J Agric Food Chem
41: 1467-1470

4. Perkins EG (1996) Volatile odor and flavor components formed in deep frying. In: Perkins
EG, Erickson MD (eds) Deep frying: chemistry, nutrition and practical applications. AOCS,
Champaign, pp 43-48

5. Stevenson SG, Jeffery L, Vaisey-Genser M, Fyfe B, Hougen FW, Eskin NAM (1984)
Performance of canola and soybean fats in extended frying. Can Inst Food Sci Technol J 17:187-
194

6. Przybylski R, Eskin NAM (1995) Methods to measure volatile compounds and the flavor
significance of volatile compounds. In: Warner K, Eskin NAM (eds) Methods to assess quality
stability of oils and fat-containing foods. AOCS, Champaign, pp 107-133

7. Melton SL, Jafar S, Sykes D, Trigiano MK (1994) Review of stability measurements for
frying oils and fried food flavor. J Am Oil Chem Soc 71:1301-1308

8. Al-Kahtani HA (1991) Survey of quality of used frying oils from restaurants. J Am Oil Chem
Soc 68:857-862

9. Che Man YB, Tan CP (1999) Effects of natural and synthetic antioxidants in refined,
bleached, and deodorized palm olein during repeated deep-fat frying of potato chips. J Am Oil
Chem Soc 76:331-339

10. Achir N, Kara W, Chipeaux C, Trezzani I, Cuvelier ME (2006) Effect of energy transfer
conditions on the chemical degradation of frying oil. Eur J Lipid Sci Technol 108:999-1006
11. Valdes AF, Garcia AB (2006) A study of the evolution of the physicochemical and structural
characteristics of olive and sunflower oils after heating at frying temperatures. Food Chem
98:214-219

12. Fiselier K, Bazzocco D, Gama-Baumgartner F, Grob K (2006) Influence of frying


temperature on acrylamide formation in French fries. Eur Food Res Technol 222:414-419

13. Chang SS, Peterson RJ, Ho CT (1978) Chemical reactions involved in deep-fat frying of
foods. J Am Oil Chem Soc 55:718-727

14. Fritsch CW (1981) Measurements of frying fat deterioration: a brief review. J Am Oil Chem
Soc 58:272-274

15. Guillen MD, Cabo N (2002) Fourier transform infrared spectra data versus peroxide and
anisidine values to determine oxidative stability of edible oils. Food Chem 77:503-510

16. Dobarganes C, Marquez-Ruiz G (2003) Oxidized fats in foods. Curr Opin Clin Nutr Metab
Care 6:157-163

17. Firestone D (1999) Official methods and recommended practices of the American Oil
Chemists' Society, 5th edn. AOCS, Champaign

18. Velasco J, Berdeaux O, Marquez-Ruiz G, Dobarganes MC (2002) Sensitive and accurate


quantification of monoepoxy fatty acids in thermooxidized oils by gas-liquid chromatography. J
Chromatogr A 982:145-152

19. Association of Official Analytical Chemists (1990) Official methods of analysis of the
association of official analytical chemists' 15th edn. AOAC Inc. Arlington, Method 982.27

20. International Organization for Standardization (2004) Animal and vegetable fats and oils-
determination of anisidine value, ISO, Geneva, Standard No. 6885

21. International Organization for Standardization (2007) Animal and vegetable fats and oils-
determination of polymerized triglycerides content by high-performance size-exclusion
chromatography (HPSEC), ISO, Geneva, Standard No. 16931

22. Marquez-Ruiz G, Jorge NI, Martfn-Polvillo M, Dobarganes MC (1996) Rapid, quantitative


determination of polar compounds in fats and oils by solid-phase extraction and size-exclusion
chromatography using monostearin as internal standard. J Chromatogr A 749:55-60

23. Orthoefer FT, Cooper DS (1996) Initial quality of frying oil. In: Perkins G, Erickson MD
(eds) Deep frying: chemistry, nutrition and practical applications. AOCS, Champaign, pp 29-42

24. White PJ (1981) Methods for measuring changes in deep-fat frying oils. Food Technol 45:75-
80
25. Firestone D, Stier RF, Blumenthal M (1991) Regulation of frying fats and oils. Food Technol
45:90-94

26. Tyagi VK, Vasishtha AK (1996) Changes in the characteristics and composition of oils
during deep-fat frying. J Am Oil Chem Soc 73:499-506

27. Houhoula DP, Oreopoulou V, Tzia C (2002) A kinetic study of oil deterioration during frying
and a comparison with heating. J Am Oil Chem Soc 79:133-137

28. Fournier V, Destaillats F, Juaneda P, Dionisi F, Lambelet P, Sebedio JL, Berdeaux O (2006)
Thermal degradation of longchain polyunsaturated fatty acids during deodorization of fish oil.
Eur J Lipid Sci Technol 108:33-42

29. Augustin MA, Asap T, Heng LK (1987) Relationships between measurements of fat
deterioration during heating and frying in RBD olein. J Am Oil Chem Soc 64:1670-1675

30. Önal B, Ergin G (2002) Antioxidative effects of Į-tocopherol and ascorbyl palmitate on
thermal oxidation of canola oil. Nahrung 46:420-426

31. Velasco J, Marmesat S, Marquez-Ruiz G, Dobarganes MC (2004) Formation of short-chain


glycerol-bound oxidation products and oxidized monomeric triacylglycerols during deep-frying
and occurrence in used frying fats. Eur J lipid Sci Technol 106:728-735

32. Peers KE, Swoboda AT (1982) Deterioration of sunflower seed oil under simulated frying
conditions and during small-scale frying of potato chips. J Sci Food Agric 33:389-395

33. Pritchard MK, Adam LR (1994) Relationship between fry color and sugar concentration in
stored Russet Burbank and Shepody potatoes. Am Potato J 71:59-68

34. Normand L, Eskin NAM, Przybylski R (2001) Effects of tocopherols on the frying stability
of regular and modified canola oils. J Am Oil Chem Soc 78:369-373

35. Lopez-Varela S, Sanchez-Muniz FJ, Garrido-Polonio C, Arroyo R, Cuesta C (1995)


Relationship between chemical and physical indexes and column HPSE chromatography
methods for evaluating frying oil. Z Ernahrungswiss 34:308-313

Anda mungkin juga menyukai