Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Gourmet and Health-Promoting Specialty Oils
Gourmet and Health-Promoting Specialty Oils
Gourmet and Health-Promoting Specialty Oils
Ebook1,149 pages12 hours

Gourmet and Health-Promoting Specialty Oils

Rating: 3 out of 5 stars

3/5

()

Read preview

About this ebook

The third volume in the AOCS PRESS MONOGRAPH SERIES ON OILSEEDS is a unique blend of information focusing on edible oils. These oils contain either unique flavor components that have lead to their being considered "gourmet oils," or contain unique health-promoting chemical components. Each chapter covers processing, edible and non-edible applications, lipids, health benefits, and more related to each type of oil.
  • Includes color illustrations of over 20 health-promoting specialty oils
  • Comprehensive resource for the chemical and physical properties and extraction and processing methods of these specialty oils
  • Describes and and includes the health effects of over 50 different oils from plants, algae, fish, and milk
LanguageEnglish
Release dateAug 25, 2015
ISBN9780128043516
Gourmet and Health-Promoting Specialty Oils

Related to Gourmet and Health-Promoting Specialty Oils

Related ebooks

Medical For You

View More

Related articles

Related categories

Reviews for Gourmet and Health-Promoting Specialty Oils

Rating: 3 out of 5 stars
3/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Gourmet and Health-Promoting Specialty Oils - Robert Moreau

    21.

    1

    Olive Oil

    Diego L. García-González, Ramón Aparicio-Ruiz and Ramón Aparicio,     Instituto de la Grasa (CSIC), Padre García Tejero, 4, 41012, Sevilla, Spain

    Introduction

    The olive tree, one of the oldest known cultivated trees, is the symbol of friendship and peace; it also plays other social and religious roles described in Greek mythology and the Old Testament. More than 1275 autochthonous olive cultivars are identified and characterized (Bartolini et al., 1998) from an ancestor that is still subject to debate, but its origin is dated circa 5000 years ago though controversy also exists about its possible geographical origin in Mesopotamia. Nevertheless, Phoenicians and Greeks were responsible for the spread of the olive tree to Western regions where they traded.

    The growth of olive oil cultivation was followed by the development of olive processing. The squeezing of olives in stone mortars, operated by early farmers, was not suitable for the increasing trade of olive oil in the Roman Empire. Romans decisively contributed to technological development with the milling crusher, which expedited the crushing process, and the wooden or iron manually activated screw press. They represented the major revolution in olive processing until Joseph Graham (1795) invented the hydraulic pressing system. The third revolution was in the second half of the twentieth century with the centrifugation system. The cost reduction of fully automated systems, which produce high-quality olive oils, the industrialization of agriculture, and the nutritional benefits of consuming olive oil abruptly increased olive oil demand and hence, olive oil production.

    Nowadays, 600 million productive olive trees grow on the Earth and are spread out on 7 million ha, with pedoclimatic conditions such as those prevailing in the Mediterranean countries that account for not less than 97% of world production. The European Union is the major producer and also consumer, and surprisingly the number one exporter and the number two importer. Olive oil is marketed in accordance with the designations of the International Olive Council trade standards (IOC. 2006). Thus, virgin olive oil must be obtained by only mechanical or other physical means under conditions, particularly thermal, which do not lead to alterations in the oil. This oil has not undergone treatment other than washing, decantation, centrifugation, and filtration. Extra-virgin olive oil (EVOO) and virgin olive oil (VOO) are different edible grades of virgin olive oil. Lampante VOO is not fit for consumption and is intended for refining or for technical purposes. Refined olive oil (ROO) is the oil refined by methods that include neutralization, decolorization with bleaching earth, and deodorization. Olive oil is an edible blend of VOO and ROO. Olive–pomace oil is obtained by solvent extraction of the olive milling by-products; its triacylglycerol composition is similar to that of VOO, but some of the nonsaponifiable compounds (e.g., waxes) may differ significantly requiring the oil to be winterized before refining. Olive–pomace oil designation means a mixture refined olive-pomace oil with virgin olive oil. The result is that olive oil is the most controlled edible oil, overseen by numerous chemical standards and regulations (Table 1.1) that protect consumers against false copies.

    Table 1.1

    Olive Oil Trade Standards.

    (1): Trans oleic fatty acids (%)

    (2): Sum of trans linoleic & linolenic fatty acids (%)

    (3): Total sterol content (mg/kg)

    (4): Erythrodiol and uvaol content (% total sterols).

    (5): Wax content: C40+C42+C44+C46 (mg/kg).

    (6): Stigmastadiene content (mg/kg)

    (7): Difference between the actual and theoretical ECN42 triacylglycerol content

    (8): Content of 2-glyceryl monopalmitate; b, C16:0 ≤ 14.0% and 2P ≤ 0.9% or C16:0 > 14:0% and 2P ≤ 1.0%;,C, C16:0 >14.0% and 2P ≤ 0.9% or C16:0 > 14:0% and 2P ≤ 1.1%.

    (9): Absorbency in ultra-violet at K232

    (10): Absorbency in ultra-violet at K270

    (11): Absorbency in ultra-violet (ΔK)

    (12): Free acidity (%m/m expressed in oleic acid)

    (13): Peroxide value (in milleq. peroxide oxygen per kg/oil)

    (14): Moisture and volatile matter (% m/m)

    (15): Trace of iron (mg/kg)

    (16): Trace of copper (mg/kg)

    (17): Insoluble impurities in light petroleum (% m/m)

    (18): Δ⁷-Stigmastenol (%)

    (19): Cholesterol (%)

    (20): Brassicasterol (%)

    (21): Campesterol (%); Camp = %Campesterol

    (22): Stigmasterol (%)

    (23): The value of β-Sitosterol is calculated as sum of: Δ⁵,²³-Stigmastadienol + Clerosterol + β-Sitosterol + Sitostanol + Δ⁵-Avenasterol + Δ⁵,²⁴-Stigmastadienol.

    (24): Organoleptic characteristics: odor & taste

    (25) Organoleptic assessment: median of defect

    (26): Organoleptic assessment: median of the fruity attribute

    (27): Organoleptic assessment: color; D, light and yellow; E, light and yellow to green; F, light, yellow to brownish yellow.

    (28): Myristic acid(% m/m methylesters)

    (29): Palmitic acid (% m/m methylesters)

    (30): Palmitoleic acid (% m/m methylesters)

    (31): Heptadecanoic acid (% m/m methylesters)

    (32): Heptadecenoic acid (% m/m methylesters)

    (33): Stearic acid (% m/m methylesters)

    (34): Oleic acid (% m/m methylesters)

    (35): Linoleic acid (% m/m methylesters)

    (36): Linolenic acid (% m/m methylesters)

    (37): Arachidic acid (% m/m methylesters)

    (38): Eicosenoic acid (% m/m methylesters)

    (39): Behenic acid (% m/m methylesters)

    (40): Lignoceric acid (% m/m methylesters)

    aIOC 2006

    Source: a, IOC 2006

    The great diversity in the chemical profiles of olive oils is mainly a consequence of the numerous olive cultivars. In analyzing the constituents of olive fruit, moisture and oil content constitute 85–90% of pulp weight, while the rest is composed of organic matter and minerals. Pulp is rich in potassium as well as in major monosaccharides (e.g., xylose, galactose, mannitol, and glucose) and small amounts of organic acids. The skin, pulp, and seed contain different lipid fractions though the major components (acylglycerols) do not contribute much to olive oil characterization, but minor compounds account for 0.5–2.0% of the olive oil composition only. They consist of a series of compounds, such as sterols, alcohols, hydrocarbons, tocopherols, phenols, pigments, et cetera, and two hundred or so aromatic compounds. The large number of chemical compounds is the basis of olive oil authentication to protect against fraudulent blends and copies (Aparicio et al., 2007) as well as the determining of the VOO sensory quality by means of volatiles and phenols (Aparicio et al., 1996). Furthermore, the chemical composition—in particular the balance between mono- and polyunsaturated fatty acids and the content of phenols, tocopherols, and carotenes—explains the cardioprotective properties of the olive oil and its effects on several other pathways, including insulin sensitivity, blood pressure, inflammatory markers, and arterial wall function, among others.

    Processing

    The primitive man who accidentally crushed fallen olives and noticed that they secreted oil was the first producer. Successive civilizations of the Mediterranean basin contributed to the technological development in olive oil processing (Di Giovacchino, 2000).

    The modern concept of the processing system includes all the processes, from the harvesting of the olives to their crushing and extraction (Fig. 1.1). Olive picking is an important operation that contributes to VOO quality since its sensory quality depends on the health and ripeness of the harvested olives; olives infested by parasites or olives collected from the ground are processed to be refined. The olives are no longer hand-picked in most of the producing countries; manual or fully automated shakers are used. The percentage of olives dropped by the shakers is increasingly higher as the harvesting is delayed. However, the olive ripeness is also a determinant variable in the olive oil quality (Aparicio & Morales, 1998), and olive oil from overripened olives is characterized by a ripe flavor and a sweet taste. This oil is less cherished by consumers than an oil characterized with a slight bitter taste and a herbaceous scent. Thus, the harvesting time is a compromise between olive oil quality and harvesting efficiency.

    Fig. 1.1 Scheme of an automatic centrifugation system.

    The picked olives should be transported to olive mills in plastic cases with holes that allow air circulation to diminish the heat in the olives placed on the bottom layer; cases can vary in capacity from 25 kg to 300 kg. After transporting the olives, the olives are stored under optimal conditions from the moment of harvesting up to their processing in the mill. The time between both events should be less than two days without risking significant changes in the oil sensory profile.

    The harvesting process does not prevent the presence of foreign materials with the olives that could be harmful to the machinery (e.g., stones) and modify the sensory attributes of the extracted olive oil. The processes of removing foreign materials and washing the olives are carried out as soon as the olives reach the olive mill. After they are weighed a sample of olives representing the whole batch is randomly taken for analytical analyses (humidity, free acidity, yield, sensory assessment). The washing step avoids the VOO being characterized with negative attributes such as muddy sediment and earthy, among others (Table 1.2).

    Table 1.2

    Main Sensory Defects Produced in the VOO Extraction Process. The Causes of the Defects and Possible Volatiles Responsible for Them

    Source: Angerosa, 2002; Morales et al. 2005.

    Cleaned and washed batches of olives are then put on a moving belt and taken to the metallic crushers. They consist of a metallic body and a high-speed rotating piece of various shapes that throws the olives against a fixed or a mobile metal grating; hammers are the most common shape although other alternatives, toothed disks, cylinders, rollers, etcetera, are available. All the pieces are made of stainless steel. The olive crushing also affects the sensory profile of the resulting olive oil. A crusher with small holes in the grating gets high extraction yield, but the resulting oil is qualified with a bitter taste. The crushing process also activates various endogenous enzymes such as polyphenoloxidase and peroxidases that promote the oxidation of phenols. Furthermore, the metallic sensory attribute, an undesirable qualifier, may be detected in VOO if the inner parts of the crusher are covered with a thin layer of iron oxide that is solubilized by free fatty acids.

    A metallic crusher can cause emulsions with a negative effect on yield. The malaxation operation diminishes this effect by merging the droplets produced by crushing the olives to increase the percentage of available oil. The malaxation process involves stirring the olive paste slowly which aids in the coalescence of small drops into large ones and favors the breakage of the unbroken cells containing oil. The malaxators are two or three cylindrical vats intandem, each one with a rotating helix with several wings that mix the paste at low speed (15–20 rpm) for 30–75 min. The vats, made of stainless steel, have double walls for circulation of heated water. Temperature and time are variables of the malaxation that are automatically controlled. The temperature of the paste should not exceed 30°C to avoid a change in oil color (from yellow-green to reddish), an increase of acidity, and destruction of the volatile compounds. An increase of the mixing time contributes to the loss of the phenolic content by oxidation. New malaxators with a controlled atmosphere of around 2% oxygen content allow the extraction of VOO with better sensory and nutritional qualities (Servili et al., 2003), if the temperature is inside the range of 25–27°C and the whole process lasts less than 60 min (Aparicio et al., 1994).

    After this step, the main constituents of olive paste are liquids (olive oil and olive mill wastewater-OMW) and solids such as small pieces of kernel and tissues. The next step is the separation of olive oil from the other constituents. Modern olive mills extract VOO by means of centrifugation systems since they allow the obtaining of high-quality olive oils with less production cost. Centrifugation is a continuous process that separates olive oil from the other materials (water and solids) by Stoke’s law which states the speed at which two nonmiscible liquids are separated under the centrifugal force. Separation is carried out inside a decanter, a cylindrical bowl with a similarly shaped screw with helical blades that gyrate at 3500–3600 rpm. The small difference between the speed at which the bowl and the inner screw rotate results in the movement of the solid (olive pomace) to one end of the centrifuge, while oil and water (the other two constituents) are moved to the other end. The addition of lukewarm water increases fluidity and helps the separation of liquid and solid phases by centrifugal force. The addition of water also increases the amount of OMW (97.2 L/100 kg olives) that has negative environmental effects due to its resistance to biodegradation. Thus, a new decanter is implemented in modern olive mills to avoid OMW contamination. It is the so-called two-phase decanter, in opposition to the three-phase decanter, that does not need the water addition.

    This new decanter does not show significant differences compared to the three-phase decanter concerning quality parameters, purity criteria, and sensory assessment (Table 1.3). The only exception is the bitterness perception due to a noticeably higher amount of phenols as a consequence of the absence of water during the centrifugation process. The main disadvantage is the production of a slurry of olive pomace (with 60–70% of water) that cannot be managed easily. Irrespective of the process used for the oil extraction, a final centrifugation of the produced VOO is needed to remove water and small solids from the oil. This process is carried out in vertical centrifuges that rotate at a high speed (6000–7000 rpm).

    Table 1.3

    Chemical and Sensory Information of the Products Resulting from the Processing of Olives by Three- and Two-Phase Decanters. Range of Values Corresponds to the Processing of Several Single Cultivars Processed by Both Decanters Simultaneously

    ¹Sensory assessment with hedonic scale of 100 mm;

    ²According to the method described by Aparicio et al. (1996)

    The whole process produces VOO, which is stored in large containers to protect against oxidation, and formation of by-products. The by-products are OMW and/or olive pomace and, the less important, twigs and leaves. Olive pomace is extracted by solvent (usually hexane) after a previous drying step. Olive pomace from two-phase decanters has to be heated at high temperature due to the moisture content, and it increases the cost and also the risk of generating polycyclic aromatic hydrocarbons (PAH). Nonedible grades of olive oil, such as lampante VOO and olive pomace oil (Table 1.1), are intended for refining. The process of refining follows all the steps applied to other edible oils to remove acid, color, and odor: degumming, neutralization, winterization, bleaching, and deodorization. Refining causes: (i) losses of tocopherols, phenols, and squalene, (ii) of sterols (≤15%), (iii) of linear alcohols from waxes, and (iv) the formation of conjugated double bonds and geometrical isomers, etcetera.

    Edible and Nonedible Applications

    The olive is a versatile foodstuff which has historical applications which include its use as an edible oil, first and foremost, and table olive; and also has other nonedible uses, such as soap making, cosmetics preparation, ointments, perfumes, lubrication, or lamp oils. Today, research on new environmentally friendly olive oil extraction systems focuses on the investigation of applications of the resulting by-products which can vary from identifying new bioactive compounds to biodiesel production (Fig. 1.2).

    Fig. 1.2 Basic scheme of the main edible and nonedible applications of olive oil.

    Edible Applications

    VOO with a rich and fruity flavor is best for uncooked dishes and for dressing salads, soups, vegetables, and cold meats. Light, refined olive oils are perfect for frying or for making more delicate dishes such as mayonnaise in which the obtrusive flavor of EVOO would be overwhelming. Each recipe prescribes which category of olive oil should be used. The highest consumption rate in terms of kilograms per person per year is found in Greece (24 kg)—followed by Spain (13.1 kg) and Italy (12.5 kg)—with a sharp increase of 5–7% per year. The United States and Japan lead the consumption increase, 12% and 30%, respectively, among nonproducer countries (Mataix & Barbancho, 2006). In fact, the world olive oil demand has increased in the last decade due to the discovery of olive oil as an everyday cooking oil by new consumers who are abandoning the idea of buying olive oil as a gourmet oil to be used in selected meals only (García-Martínez, et al. 2002). However, standard olive oil needs to be repositioned and distanced from EVOO by giving customers enough cues to revise their opinions and consumption strategies. Furthermore, the awareness of consumers with a higher sensory quality and a clear geographical identity has encouraged producers to develop new products based on varietal VOOs, characterized by different sensory attributes and VOOs with a declared geographical origin (Protected Designation of Origin or PDO) that are controlled by a Regulatory Council when produced within the European Union. Today, olive oil authenticity, sold by PDOs and reputed sellers, is guaranteed, and the consumers’ expectations of sensory quality are fulfilled when they buy a VOO bottle.

    ROO and even VOO are used in fish canning (e.g., anchovy and tuna) because this combines the positive qualities of ω-3 polyunsaturated fatty acids and oleic acid from the coverage oil, which makes the result ideal for preventing cardiovascular diseases. In the case of VOO, a partitioning also exists toward the brine phase of major phenol compounds with all the advantages of these antioxidant compounds. On the other hand, producers recently released new products based on virgin or olive oil flavored with spices and herbs, from garlic to rosemary, and occasionally mixed with vinegar, ready to be used as a salad dressing or dip.

    Deep- and pan-frying are the second most important applications of olive oil, an ancient culinary practice of the Mediterranean countries. The composition of olive oil, with a low content of saturated fatty acid and a high content of antioxidants (e.g., phenols), makes it an excellent oil to be applied in culinary processes that apply high temperatures, such as in frying and baking. After 10 hours of heating at 180°C, the total polar compounds of VOO can be lower than 25%, while other vegetable oils, such as sunflower oil or commercial blends intended for frying purposes, easily reach 29% of polar compounds (Kalantzakis et al., 2006). Furthermore, some varietal VOO (e.g., Picual) are even more resistant to thermoxidation, and after 25 hours of heating at 180°C, they only contain 15% of polar compounds (Brenes et al., 2002). Although some phenols (e.g., hydroxytyrosol and tyrosol-like substances) are dramatically decreased in the first hours of frying, other phenols (1-acetoxypinoresinol and pinoresinol) remain in the oil after 25 hours. Furthermore, high-stability olive oil has a low melting point that allows it to easily drain from fried foodstuffs, the other property appreciated in frying oils.

    Nonedible Applications

    Nonedible applications of olive oil have existed since ancient times, and some of them remain today. Thus, olive oil is used in such preparations as lip balm, shampoo, and hand lotion, as well as in dry skin and dandruff treatments. Furthermore, olive oil contains antioxidant and anti-inflammatory compounds that make it suitable for topical applications in skin damage and dermatitis. Together with the traditional nonedible uses of olive oil, the reuse of by-products is being investigated to avoid environmental problems and also to reduce the high cost of olive oil production, thus increasing the competitiveness. The by-products are mainly obtained from olive cake, olive stones, and OMW by means of pyrolysis, combustion, and gasification processes, among others (Paraskeva & Diamadopoulos, 2006).

    The olive cake can be an energy source: one can substitute 1100 tons of exhausted olive cake for 420 tons of No. 2 heavy fuel containing 4% sulfur (Masghouni & Hassairi, 2000). The use of this cake as an energy source is environmently friendly and avoids the emissions of sulfur. The cake can also be used for animal feeding or returned to the olive trees as mulch. Furthemore, the olive husk is used in molded products, plastics, in furfural manufacturing, and in activated carbon production, as well as fuel. OMW is also used as a growth medium for microbial production of enzymes (D’Annibale et al., 2006), algal biomass rich in polyunsaturated fatty acids (Anon., 2004), and for clay brick manufacturing (Mekki et al., 2006); its flocculated solid content is used as soil organic fertilizer (García-GÓmez et al., 2003). Furthermore, various compounds are extracted from OMW and other by-products to be used as food additives, cosmetics, and medicines; OMW is, for example, a source of antioxidants as it contains approximately 53% of the olive phenols. Research efforts were recently made to develop cost-effective approaches to efficiently recover phytosterols from OMW and leaves, and to obtain pure crystals from the resulting sterol concentrate slurries.

    Acyl Lipids and Fatty Acids Composition

    Lipid Biosynthesis in Olives

    Fatty acids account for up to 85% of olive oil total composition in the form of triacylglycerols (TAG). Thus, in discussing olive oil composition, one should first consider the biosynthesis of fatty acids. Labeling experiments demonstrated the existence in the pulp (pericarp) of developing olives of two carboxylating mechanisms —both in the light and dark—that involve the activity of three enzymes: ribulose-1,5-bisphosphate carboxylase, phosphoenolpyruvirate carboxylase, and NAD-malate dehydrogenase (Sánchez & Salas, 1997). Studies under autotrophic and heterotrophic conditions indicated that oil formation in olives requires the contribution of the two sources of reduced carbon, leaf, and fruit photosynthesis. Furthermore, experiments using either radio labeled acetate or pyruvate showed that both precursors are efficiently incorporated into acyl lipids by tissue slices from developing olives, suggesting that two operative pathways are present in the olive pulp for the formation of acetyl-CoA, the precursor for fatty acid biosynthesis. One pathway is based on the breakdown of six carbon sugars via glycolysis in the plastid, while the other involves collaboration with mitochondria (Harwood & Sánchez, 2000).

    De novo fatty acid biosynthesis in olive fruit, as in most oil crops, involves the concerted activity of two enzymes, acetyl-CoA carboxylase and fatty acid synthase, that catalyze multiple reactions. In olive fruit, two isoforms of acetyl-CoA carboxylase exist, the plastid-localized isoform being a multienzyme complex. The fact that dicotyledons retain a multifunctional protein in the epithelial cytosol is important in providing malonyl-CoA for fatty acid elongation. Once malonyl-CoA is generated, it is used to form malonyl-ACP for fatty acid synthesis whose reactions comprise seven steps. Sánchez & Harwood (1992) studied the overall fatty acid synthase activity in soluble fractions of olive pulp by using radiolabeled malonyl-CoA as the precursor. They concluded that the activity was stimulated by ACP and inhibited by cerulenin, and it showed a strong dependence on NADH, NADPH, and thiol reagents. Chain termination of fatty acid synthesis is achieved through the activity of acyl-ACP thioesterase that shows better activity with C18 (oleoyl)-ACP. Further desaturations to produce polyunsaturated fatty acids occur either within the plastid or on the endoplasmic reticulum. Those fatty acids produced in the plastid are channeled out of the organelle to form the acyl-CoA pool in the cytosol. Further desaturation of oleate to linoleate and then to α-linoleate take place via phosphatidylcholine substrates on the endoplasmic reticulum.

    The incorporation of fatty acids into complex lipids occurs via the well-known Kennedy pathway, which has a series of four reactions that yield triacyglycerols as the end product. In the case of olives, the incorporation of acyl-CoAs into TAGs by microsomes is mediated by the initial incorporation into phosphatidylcholine; oleoyl-CoA was found to be a better substrate for glycerolipid synthesis than palmitoyl-CoA (Sánchez et al., 1992). Furthermore, stereospecific analyses of olive oil showed that linoleonyl-CoA is also an effective substrate for olive G3PAT, while in vivo labeling experiments, using 14C-acetate and tissue slices, showed TAG formation is strongly reduced at temperatures above 40°C with a commensurate rise in the relative labeling of diacylglycerols (Rutter et al., 1997) that might be indicative of the limiting activity of DAGAT at high temperature.

    Fleshy pericarp accumulates the vast majority of lipids of the olive, but TAGs are also stored within the seeds in oil bodies. Ultrastructural studies of olive pulp showed that no such oil bodies are observed in mature fruits, as they fuse on contact to form oil droplets during the fruit maturation. In fact, new oil bodies appear to form then, and they fuse with oil droplets to result in the formation of large oil

    Enjoying the preview?
    Page 1 of 1