Anda di halaman 1dari 9

Pag e 299

patterns of products formed are much simpler than those obtained by heating at frying temperatures (180C) for 1 hr.
5.5.6 Concluding Thoughts on Lipid Oxidation in Food
We are witnessing a revolution in the field of lipid oxidation caused by remarkable advances in the sciences, analytical
methodology, and cross-disciplinary communication. Some of the long established theories, which in the past helped us to
develop an understanding of the basic mechanisms of lipid oxidation, must now give way to new thinking and a new approach to
the examination and control of oxidizing lipids in food.
It may not be too unreasonable to envision lipid oxidation in food in the same way one looks at the weather map in the United
States. The atmospheric phenomena and weather activities vary from region to region and the weather map may change
significantly from moment to moment:
1. Lipid oxidation is not a single process that proceeds according to a strict sequence. Rather, numerous events take place in
different regions of the food matrix. The different reactions usually occur with considerable overlap, and the rate of each may
vary depending on conditions in each region.
2. The conditions in each region (oxidation parameters) are not static. They change continuously.
3. In food, oxidation involves oxidizable lipid substrates differing in composition, chemical and physical properties, and in their
sensitivity to oxidation (fatty acids, mixtures of acylglycerols, phospholipids).
4. The lipid substrates coexist in proximity with various other nonlipid, major (proteins, carbohydrates, water) and minor (trace
metals, vitamins, enzymes, pro- and antioxidants) components. A multitude of interactions take place.
5. Nearly all oxidative events are not strictly chemical. Instead, they are largely dictated by the physical state of both substrate
and medium.
6. The minor components play a major role in a critical, delicate oxidative-antioxidative balance, the mechanism of which is not
fully understood.
7. The same inhibitors (or accelerators) may play opposite roles depending on the oxidative event being considered. For
example, a phenolic antioxidant may block the formation of volatile oxidation products while enhancing the accumulation of
peroxides.
It is clear that a realistic evaluation of lipid oxidation in food can only be achieved if the foregoing scenario, in all of its detail, is
considered.
5.6 Chemistry of Fat and Oil Processing
5.6.1 Refining
Crude oils and fats contain varying amounts of substances that may impart undesirable flavor, color, or keeping quality. These
substances include free fatty acids, phospholipids, carbohydrates, proteins and their degradation products, water, pigments
(mainly carotenoids and chlorophyll), and fat oxidation products. Crude oils are subjected to several commercial refining
processes designed to remove these materials.

Pag e 300

5.6.1.1 Settling and Degumming


Settling involves heating the fat and allowing it to stand until the aqueous phase separates and can be withdrawn. This rids the fat
of water, proteinaceous material, phospholipids, and carbohydrates. In some cases, particularly with oils containing substantial
amounts of phospholipids (e.g., soybean oil), a preliminary treatment known as degumming is applied by adding 23% water,
agitating the mixture at about 50C, and separating the hydrated phospholipids by settling or centrifugation.
5.6.1.2 Neutralization
To remove free fatty acids, caustic soda in the appropriate amounts and strength is mixed with the heated fat and the mixture is
allowed to stand until the aqueous phase settles. The resulting aqueous solution, called foots or soapstock, is separated and used
for making soap. Residual soapstock is removed from the neutral oil by washing it with hot water, followed by settling or
centrifugation.
Although free fatty acid removal is the main purpose of the alkali treatment, this process also results in a significant reduction of
phospholipids and coloring matter.
5.6.1.3 Bleaching
An almost complete removal of coloring materials can be accomplished by heating the oil to about 85C and treating it with
adsorbants, such as fuller's earth or activated carbon. Precautions should be taken to avoid oxidation during bleaching. Other
materials, such as phospholipids, soaps, and some oxidation products, are also absorbed along with the pigments. The bleaching
earth is then removed by filtration.
5.6.1.4 Deodorization
Volatile compounds with undesirable flavors, mostly arising from oxidation of the oil, are removed by steam distillation under
reduced pressure. Citric acid is often added to sequester traces of pro-oxidant metals. It is believed that this treatment also
results in thermal destruction of nonvolatile off-flavor substances, and that the resulting volatiles are distilled away.
Although the oxidative stability of oils is generally improved by refining, this is not always the case. Crude cottonseed oil, for
example, has a greater resistance to oxidation than its refined counterparts due to the greater amounts of gossypol and
tocopherols in the crude oil. On the other hand, there can be little doubt as to the remarkable quality benefits that accrue from
refining edible oils. An impressive example is the upgrading of palm oil quality that has occurred in recent years. Furthermore, in
addition to the obvious improvements in color, flavor, and stability, powerful toxicants (e.g., aflatoxins in peanut oil and gossypol
in cottonseed oil) are effectively eliminated during the refining process.
5.6.2 Hydrogenation
Hydrogenation of fats involves the addition of hydrogen to double bonds in the fatty acid chains. The process is of major
importance in the fats and oils industry. It accomplishes two major objectives. First, it allows the conversion of liquid oils into
semisolid or plastic fats more suitable for specific applications, such as in shortenings and margarine, and second, it improves the
oxidative stability of the oil.
In practice, the oil is first mixed with a suitable catalyst (usually nickel), heated to the desired temperature (140-225C), then
exposed, while stirred, to hydrogen at pressures up to 60 psig. Agitation is necessary to aid in dissolving the hydrogen, to
achieve uniform mixing of

Pag e 301

the catalyst with oil, and to help dissipate the heat of the reaction. The starting oil must be refined, bleached, low in soap, and
dry; the hydrogen gas must be dry and free of sulfur, CO2, or ammonia; and the catalyst must exhibit long-term activity, function
in the desired manner with respect to selectivity of hydrogenation and isomer formation, and be easily removable by filtration.
The course of the hydrogenation reaction is usually monitored by determining the change in refractive index, which is related to
the degree of saturation of the oil. When the desired end point is reached, the hydrogenated oil is cooled and the catalyst is
removed by filtration.
5.6.2.1 Selectivity
During hydrogenation, not only are some of the double bonds saturated, but some may also be relocated and/or transformed
from the usual cis to the trans configuration. The isomers produced are commonly called iso acids. Partial hydrogenation thus
may result in the formation of a relatively complex mixture of reaction products, depending on which of the double bonds are
hydrogenated, the type and degree of isomerization, and the relative rates of these various reactions. A simplistic scheme
showing the possible reactions that linolenate can undergo during hydrogenation is shown here:

In the case of natural fats, the situation is further complicated by the fact that they already contain an extremely complex mixture
of starting materials.
The term selectivity refers to the relative rate of hydrogenation of the more unsaturated fatty acids as compared with that of the
less unsaturated acids. When expressed as a ratio (selectivity ratio), a quantitative measure of selectivity can be obtained in more
absolute terms. The term selectivity ratio, as defined by Albright [3], is simply the ratio (rate of hydrogenation of linoleic to
oleic)/(rate of hydrogenation of oleic to stearic). Reaction rate constants can be calculated from the starting and ending fatty acid
compositions and the hydrogenation time (Fig. 35). For the reactions just mentioned, the selectivity ratio (SR) is

FIGURE 35
Reaction rate constants for hydrog enation of soybean oil.
(From Ref. 5.)

Pag e 302

FIGURE 36
SR curves for soybean oil [3],
IV +
decrease in
iodine value; L/L 0 + fraction of linoleic
acid
unhydrog enated.

K2/K3 = 0.159/0.013 = 12.2, which means that linoleic acid is being hydrogenated 12.2 times faster than oleic acid.
Since calculations of SR for every oil hydrogenated would be quite tedious, Albright [3] prepared a series of graphs for various
oils by calculation of the fatty acid composition at constant SR. In these graphs the decrease in iodine value (IV) is plotted
against the fraction of linoleic acid that remains unhydrogenated (L/L0). Although the curves are calculated with the assumption
that the reaction rates are first order and that isooleic acid is hydrogenated at the same rate as oleic, they are nonetheless very
useful for predicting selectivity. Selectivity ratio curves for soybean oil (K2/K3) are shown in Figure 36. Of course, linolenic acid
selectivity can be similarly expressed; that is, LnSR = K1/K2 where K1 and K2 are as defined in Figure 30. This is relevant to the
hydrogenation of soybean oil, since flavor reversion in this oil is believed to arise from its linolenate content.
Different catalysts result in different selectivities, and operating parameters also have a profound effect on selectivity. As shown
in Table 8, larger SR values result from high temperatures, low pressures, high catalyst concentration, and low intensity of
agitation. The effects of processing conditions on the rate of hydrogenation and on the formation of trans acids are also
TABLE 8 Effects of Processing Parameters on Selectivity and Rate of
Hydrog enation
SR

Trans acids

Rate

Hig h temperature

Hig h

Hig h

Hig h

Hig h pressure

Low

Low

Hig h

Hig h catalyst concentration

Hig h

Hig h

Hig h

Hig h-intensity ag itation

Low

Low

Hig h

Processing parameter

Pag e 303

given. Several mechanistic speculations have been advanced to explain the observed influence of process conditions on
selectivity and rate of hydrogenation, and these are discussed next.
5.6.2.2 Mechanism
The mechanism involved in fat hydrogenation is believed to be the reaction between unsaturated liquid oil and atomic hydrogen
adsorbed on a metal catalyst. First, a carbon-metal complex is formed at either end of the olefinic bond (complex a in Fig. 37).
This intermediate complex is then able to react with an atom of catalyst-adsorbed hydrogen to form an unstable halfhydrogenated state (b or c in Fig. 37) in which the olefin is attached to the catalyst by only one link and is thus free to rotate. The
half-hydrogenated compound now may either (a) react with another hydrogen atom and dissociate from the catalyst to yield the
saturated product (d in Fig. 37), or (b) lose a hydrogen atom to the nickel catalyst to restore the double bond. The regenerated
double bond can be either in the same position as in the unhydrogenated compound, or a positional and/or geometric isomer of
the original double bond (e and f in Fig. 37).
In general, evidence seems to indicate that the concentration of hydrogen adsorbed on the catalyst is the factor that determines
selectivity and isomer formation [5]. If the catalyst is saturated with hydrogen, most of the active sites hold hydrogen atoms and
the chance is greater that two atoms are in the appropriate position to react with any double bond upon approach. This results in
low selectivity, since the tendency will be toward saturation of any double bond approaching the two hydrogens. On the other
hand, if the hydrogen atoms on the catalyst are scarce, it is more likely that only one hydrogen atom reacts with the double
bonds, leading to the half-hydrogenation-dehydrogenation sequence and a greater likelihood of isomerization. Thus, operating
conditions (hydrogen pressure, intensity of agitation, temperature, and kind and concentration of catalyst) influence selectivity
through their effect on the ratio of hydrogen to catalyst sites. An increase in temperature, for example, increases the speed of the
reaction and causes faster removal of hydrogen from the catalyst, giving rise to increased selectivity.
The ability to change the SR by changing the processing conditions enables processors to exert considerable control over the
properties of the final oil. A more selective hydrogenation, for example, allows linoleic acid to be decreased and stability to be
improved while minimizing

FIGURE 37
Half-hydrog enation-hydrog enation reaction scheme.
Asterisk indicates metal link.

Pag e 304

the formation of fully saturated compounds and avoiding excessive hardness. On the other hand, the more selective the reaction,
the greater will be the formation of trans isomers, which are of concern from a nutritional standpoint. For many years
manufacturers of food fats have been trying to devise hydrogenation processes that minimize isomerization while avoiding the
formation of excessive amounts of fully saturated material.
5.6.2.3 Catalysts
As indicated earlier, catalysts vary with regard to the degree of selectivity they provide. Nickel on various supports is almost
invariably used commercially to hydrogenate fats. Other catalysts, however, are available. These include copper,
copper/chromium combinations, and platinum. Palladium has been found to be considerably more efficient (in terms of the
amount of catalyst required) than nickel, although it produces a high proportion of trans isomers. The so-called homogenous
catalysts, which are soluble in oil, provide greater contact between oil and catalyst and better control of selectivity. A host of
different compounds are capable of poisoning the catalyst used, and these compounds are often the major source of problems
encountered during commercial hydrogenation. Poisons include phospholipids, water, sulfur compounds, soaps, certain glycerol
esters, CO2, and mineral acids.
5.6.3 Interesterification
It has been mentioned that natural fats do not contain a random distribution of fatty acids among the glyceride molecules. The
tendency of certain acids to be more concentrated at specific sn positions varies from one species to another and is influenced
by factors such as environment and location in the plant or animal. The physical characteristics of a fat are greatly affected not
only by the nature of constituent fatty acids (i.e., chain length and unsaturation) but also by their distribution in the triacylglycerol
molecules. Indeed, unique fatty acid distribution patterns of some natural fats limit their industrial applications. Interesterification
is one of the processes that can be applied to improve the consistency of such fats and to improve their usefulness. This process
involves rearranging the fatty acids so they become distributed randomly among the triacylglycerol molecules of the fat.
5.6.3.1 Principle
The term interesterification refers to the exchange of acyl radicals between an ester and an acid (acidolysis), or an ester and an
alcohol (alcoholysis), or an ester and an ester (transesterfication). It is the latter reaction that is relevant to industrial
interestification of fat, also known as randomization, since it involves ester interchange within a single triacylglycerol molecule
(intraesterification) as well as ester exchange among different molecules.
If a fat contains only two fatty acids (A and B), eight possible triacylglycerol species n3 are possible according to the rule of
chance:

Regardless of the distribution of the two acids in the original fat (e.g., AAA and BBB or ABB, ABA, BBA), interesterification
results in the shuffling of fatty acids within a single molecule and among triacylglycerol molecules until an equilibrium is achieved
in which all possible combinations are formed. The quantitative proportions of the different species depend on the

Pag e 305

amount of each acid in the original fat and can be predicted by simple calculation, as already discussed under the 1,2,3-random
distribution hypothesis.
5.6.3.2 Industrial Process
Interesterification can be accomplished by heating the fat at relatively high temperatures (<200C) for a long period. However,
catalysts are commonly used that allow the reaction to be completed in a short time (e.g., 30 min) at temperatures as low as
50C. Alkali metals and alkali metal alkylates are effective low-temperature catalysts, with sodium methoxide being the most
popular. Approximately 0.1% catalyst is required. Higher concentrations may cause excessive losses of oil resulting from the
formation of soap and methyl esters.
The oil to be esterified must be extremely dry and low in free fatty acids, peroxides, and any other material that may react with
sodium methoxide. Minutes after the catalyst is added the oil acquires a reddish brown color due to the formation of a complex
between the sodium and the glycerides. This complex is believed to be the true catalyst. After esterification, the catalyst is
inactivated by addition of water or acid and removed.
5.6.3.3 Mechanisms
Two mechanisms for interesterification have been proposed [135].
Enolate Ion Formation
According to this mechanism, an enolate ion (II), typical of the action of a base on an ester, is formed. The enolate ion reacts
with another ester group in the triacylglycerol molecule to produce a -keto ester (III), which in turn reacts further to give
another -keto ester (IV). Intermediate IV yields the intramolecularly esterified product V.

The same mode of action applies to ester interchange between two or more triacylglycerol molecules. The intra-ester-ester
interchange is believed to predominate in the initial stages of the reaction.
Carbonyl Addition
In this proposed mechanism, the alkylate ion adds onto a polarized ester carboxyl, producing a diglycerinate intermediate:

Pag e 306

This intermediate reacts with another glyceride by abstracting a fatty acid, thus forming a new triacylglycerol and regenerating a
diglycerinate for further reaction. Ester interchange between fully saturated S3 and unsaturated U3 molecules is shown here:

5.6.3.4 Directed Interesterification


A random distribution, such as that produced by intersterification, is not always the most desirable. Interesterification can be
directed away from randomness if the fat is maintained at a temperature below its melting point. This results in selective
crystallization of the trisaturated glycerides, which has the effect of removing them from the reaction mixture and changing the
fatty acid equilibrium in the liquid phase. Interesterification proceeds with the formation of more trisaturated glycerides than
would have otherwise occurred. The newly formed trisaturated acylglycerols crystallize and precipitate, thus allowing the
formation of still more trisaturated glycerides, and the process continues until most of the saturated fatty acids in the fat have
precipitated. If the original fat is a liquid oil containing a substantial amount of saturated acids, it is possible, by this method, to
convert the oil into a product with the consistency of shortening without resorting to hydrogenation or blending with a hard fat.
The procedure is relatively slow due to the low temperature used, the time required for crystallization, and the tendency of the
catalyst to become coated. A dispersion of liquid sodium-potassium alloy is commonly used to slough off the coating as it forms.
Rearrangement can also be selectively controlled during interesterification by adding excess fatty acids and continuously distilling
out the liberated acids that are highly volatile. This impoverishes the fat of its acids of lower molecular weight. The content of
certain acids in a fat also can be reduced by using suitable solvents to extract appropriate acids during the interesterification
process.

Pag e 307

FIGURE 38
Effect of interesterification on solid
content index. (From Ref. 135.)

5.6.3.5 Applications
Interesterification finds its greatest application in the manufacture of shortenings. Lard, due to its high proportion of disaturated
triacylglycerols with palmitic acid in the 2 position, forms relatively large and coarse crystals, even when rapidly solidified in
commercial chilling machines. Shortenings made from natural lard possess a grainy consistency and exhibit poor performance in
baking. Randomization of lard improves its plastic range and makes it a better shortening. Directed interesterification, however,
produces a product with a higher solids content at high temperatures (Fig. 38) and thus extends its plastic range.
Salad oil of a relatively low cloud point can be made from palm oil by fractionation after directed interesterification. The use of
interesterification has also been applied to the production of high-stability margarine blends and hard butters that have highly
desirable melting qualities.
Using a column with countercurrent flow of dimethylformamide, a process has been developed to selectively reduce the content
of linolenic acid in soybean oil by direct interesterification.
5.7 Role of Food Lipids in Flavor
5.7.1 Physical Effects
Pure food lipids are nearly odorless. However, apart from their major contributions as precursors of flavor compounds, they
modify the overall flavor of many foods through their effect on mouth feel (e.g., the richness of whole milk and the smooth or
creamy nature of ice cream) and on the volatility and threshold value of the flavor components present.

Anda mungkin juga menyukai