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

Only $11.99/month after trial. Cancel anytime.

Food Science and the Culinary Arts
Food Science and the Culinary Arts
Food Science and the Culinary Arts
Ebook1,190 pages17 hours

Food Science and the Culinary Arts

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Food Science and the Culinary Arts is a unique reference that incorporates the principles of food and beverage science with practical applications in food preparation and product development. The first part of the book covers the various elements of the chemical processes that occur in the development of food products. It includes exploration of sensory elements, chemistry, and the transfer of energy and heat within the kitchen. The second part looks in detail at the makeup of specific foodstuffs from a scientific perspective, with chapters on meat, fish, vegetables, sugars, chocolate, coffee, and wine and spirits, among others. It provides a complete overview of the food science relevant to culinary students and professionals training to work in the food industry.

  • Provides foundational food science information to culinary students and specialists
  • Integrates principles of food science into practical applications
  • Spans food chemistry to ingredients, whole foods, and baked and mixed foods
  • Includes a comprehensive glossary of terms in food science
LanguageEnglish
Release dateJan 4, 2018
ISBN9780128118177
Food Science and the Culinary Arts
Author

Mark Gibson

Mark Gibson has a broad and varied professional background in healthcare, and is both a mental health and general nurse. His experience includes working in general and clinical management as well as higher education, with his specialist areas of interest being that of healthcare governance, patient safety and incident investigation, legal matters and staff training and development.

Related to Food Science and the Culinary Arts

Related ebooks

Food Science For You

View More

Related articles

Reviews for Food Science and the Culinary Arts

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Food Science and the Culinary Arts - Mark Gibson

    inspiration.

    Part 1

    Science Knowledge and Discipline

    Chapter 1

    Food Groups

    Abstract

    Many people might be surprised to learn that much of the food we eat is still alive, whether raw oysters, uncooked fruit and vegetables, fresh seeds, nuts and herbs, and fermented and bacterially enhanced foods like yogurts, cheeses, sauerkraut, miso, and tempeh. As a result, this chapter takes a look at the various foods and the nutritional components that make up our food requirements, not only at macronutrient levels but also at the micronutrient level too. This is summarized in six categories: carbohydrates, proteins, fats, water, vitamins and minerals. The chapter also delves into the various metabolic processes whereby food is metabolized by the body for use in its many forms. A quick look at the body's energy preferences and the chapter concludes with recommended energy balance portfolios.

    Keywords

    Macronutrients; Carbohydrates; Proteins; Fats; Vitamins; Minerals; Metabolism; Energy portfolio

    Chapter Outline

    1.1Macronutrients

    1.1.1Carbohydrates

    1.1.2Proteins

    1.1.3Fats

    1.2Water

    1.3Micronutrients

    1.3.1Vitamins

    1.3.2Minerals

    1.3.3A Good Balanced Energy Portfolio

    1.4Metabolism

    1.5Energy Preferences

    1.5.1Carbohydrate Metabolism

    1.5.2Lipid (Fat) Metabolism

    1.5.3Protein Metabolism

    References

    Further Reading

    Many might be surprised to learn that much of the food we eat is still alive, whether raw oysters, uncooked fruit and vegetables, fresh seeds, nuts and herbs, and fermented and bacterially enhanced foods like yogurts, cheeses, sauerkraut, miso, and tempeh. They are alive in the sense that they contain cells that are still living (Field, 2011). The human body then, benefits from both raw and cooked foods. It must also be remembered that of the many foods we eat, we are never eating just one group, say vitamins (unless one is taking supplements), for example. Instead, an apple might contain several groups of foods including vitamins, minerals, fibers, and carbohydrates perhaps. It should be noted too that while certain food groups are ingested every day (some nutrients are required on a regular basis) not every food groups (in particular the micronutrients) are required on a daily basis, as certain of these can be stored in the body (Belitz et al., 2009). It is similarly noteworthy that some of these nutrients cannot be synthesized by the body and must be ingested through the food we eat.

    Before we get into the mix so to speak, it is perhaps worth taking a quick look at the various nutritional components that make up our food requirements. The term food applies to a variety of organic materials and can be categorized in a number of ways. Foods can be grouped by whether or not they provide energy (calorific value) for the body or not or more descriptively by such terms as fish, meat, dairy, fruit, and vegetables. In this book, we categorize foods according to the body's needs. As such, our first foray into the vast area of food shall take the form of a quick look, not only at the nutritional profile of macronutrient food groups but also those of the loosely termed minor (although no less important) micronutrient food groups too. These are summarized in six categories: carbohydrates, proteins, fats, water (which comprise 99% of the dry weight of the food we eat), and vitamins and minerals (collectively making up the remaining 1%).

    1.1 Macronutrients

    Macronutrients are the nutrients needed in greater quantities than those of micronutrients; these are the proteins, lipids (fats), carbohydrates, and water, which, with the exception of water, provide energy to the body at the following rates. Proteins provide around 4 Cal/g, lipids (fats) roughly 9 Cal/g, and carbohydrates supply approximately 4 Cal/g. As for water, there are no calories for energy; however, as will be seen later, it becomes clear that water is a vital and necessary component of the macronutrients (Table 1.1 and Section 1.2) (EUFIC, 2009; Wilson, 2008; WHO, 1997, 2006).

    Table 1.1

    Source: Compiled from multiple sources: Latham, M.C., 1997. Human Nutrition in the Developing World. Food and Nutrition Series—No. 29. Food and Agriculture Organisation of the United Nations, Rome; Gibney, M.J., et al., 2009. Introduction to Human Nutrition. Blackwell Publishing, Oxford; Eduweb, 2010. Fact Sheets: Body Composition. Retrieved 15 Oct 2017 from http://www.rowett.ac.uk/edu_web/; EUFIC, 1998. The European Food Information Centre: What Do We Mean by Nutrition?. Retrieved 11 November 2016, from http://www.eufic.org/article/en/health-and-lifestyle/healthy-eating/artid/nutrition-2/; Gibson, M., 2012. The Feeding of Nations: Re-defining Food Security for the 21st Century. CRC Press, Boca Raton, FL.

    a Extracellular water.

    b Intracellular water.

    1.1.1 Carbohydrates

    The basic building blocks of carbohydrates are the amalgamation of carbon, hydrogen, and oxygen. Carbohydrates are also ketones, alcohols, polyhydroxy aldehydes, and acids and their derivatives. In more lay terms, carbohydrates are one of the main macronutrients within our diets; they, along with fats and proteins, supply valuable energy for the body to function. Although in terms of ranking and as will be discussed later, carbohydrates are the most important of the energy providers (Eliasson, 2016). This is partly because our digestive system readily metabolizes carbohydrates into glucose (blood sugar) that in turn is used as energy for cells, tissues, and organs. Carbohydrate sugars can also be assembled according to their structural complexity, that is, three principal groups: the monosaccharides including glucose, fructose, and galactose; the disaccharides comprising two sugar units known as sucrose and lactose; and the polysaccharides that cover the starches, cellulose (fibers), and hemicellulose (Lee et al., 2014; Cummings and Stephen, 2007).

    Food, because of its heterogeneity, means that carbohydrates are not consumed in isolation. Instead, they are found in a variety of foods that also bring an assortment of other nutrients to the diet; these include vitamins and minerals, dietary fiber, antioxidants, and phytochemicals. These are found in varying amounts in fruits, vegetables, grain-based foods, and several dairy products.

    1.1.1.1 Monosaccharides—Simple sugars

    Monosaccharides are the simplest units of carbohydrates and the simplest forms of sugar there is. The main monosaccharides are the hexoses (simple sugars in which the molecules contain six carbon atoms)—these include glucose (known also as dextrose), fructose¹ (commonly called levulose), galactose, and mannose (Eliasson, 2016). All of which are naturally found in fruit juices. These simple sugar forms are typically colorless and crystalline water-soluble solids. Monosaccharides have a sweet taste, the best examples being fructose and galactose and to a lesser degree, glucose. Glucose, however, is without doubt the most common of the monosaccharides and is also incidentally, industrially manufactured on a large scale through the hydrolysis² of cornstarch. Being less sweet though, glucose is often used instead of table sugar (the common name for the disaccharide—sucrose) in cooking, especially in the pastry kitchen. Moreover, glucose can also be prepared in relatively pure form through hydrolysis of both disaccharides and polysaccharides. Importantly, the simple monosaccharide glucose is also the easiest sugar for the body to break down and metabolize. In fact, it is only in its simplest form, that is, glucose, that sugar can be absorbed into the blood and in a relatively short space of time too (Lee et al., 2014). This is why eating sugary foods and soft drinks very quickly causes spikes in the glucose levels of the blood, sometimes referred to as a sugar rush. Although useful in moderation, if sugar intake is continually excessive, then in the long run this can lead to heart and vascular diseases (Cummings and Stephen, 2007; Eliasson, 2016). Fructose, another simple sugar, is also found naturally in fruits, and if cane or sugar beet is hydrolyzed, one molecule of each, fructose and glucose, is created from each molecule of sucrose (see Sections 6.5 and 10.1.7.1) (Lee et al., 2014). It is also worth noting that while hydrolysis of di- or polysaccharides produces monosaccharide, monosaccharides in turn cannot be further hydrolyzed to make simpler compounds. Instead, monosaccharides in turn become the building blocks of disaccharides like sucrose (table sugar) and lactose. Sugars are used as sweeteners to improve the palatability and preservation of many foods. They are also responsible for certain functional characteristics in foods like viscosity, texture, body, and browning or caramelization properties. Sugars also increase dough yield, influence starch and protein breakdown, and control moisture, consequently preventing the drying out of certain foods (Cummings and Stephen, 2007).

    1.1.1.2 Disaccharides—Sugars

    When it comes to disaccharides like sucrose, maltose, and lactose, they may be thought of as comprising two simple sugar molecules that come together to make one larger molecule. Sucrose is the common sugar in the cook's kitchen and at home. It is made largely from sugarcane and sugar beets. The hydrolysis of sucrose results in the formation of glucose and fructose. Another monosaccharide, maltose, is the resulting compound from the hydrolysis of starch originating from a wide variety of sources (Eliasson, 2016). In turn, maltose can be hydrolyzed into two molecules of glucose. In the alcoholic fermentation process too, maltose is the resultant product from the enzymatic³ breakdown of starch from malted barley. The sugar then feeds the yeast producing alcohol and carbon dioxide. Lactose, the third disaccharide, is more commonly known as milk sugar. It can be hydrolyzed to form glucose and galactose—both simple sugars, although if they are then oxidized, both galactose and glucose form acid compounds.

    1.1.1.3 Polysaccharides—Complex sugars

    Although polysaccharides are condensation products⁴ of both hexoses (simple sugars) and other monosaccharides; three polysaccharides in particular, starch, cellulose, and hemicellulose, are of interest in the food and organic chemistry industries. Although they form part of the saccharides, one might expect polysaccharides to share the same characteristic sweetness. Instead, starch, cellulose, and hemicellulose are predominantly (but not always) insoluble and have such complex molecular structures so as to be completely dissimilar from mono- and disaccharides (Eliasson, 2016). Sometimes, these polysaccharides go by the collective name of dietary fibers. Fiber or dietary fiber has many different meanings in the world of nutrition. It is not a reference to a precise chemical component; instead, it is essentially a physiological concept that describes the proportion of food (derived from cellular walls of plants) that is ultimately poorly digested in human beings (Cummings and Stephen, 2007). Nondigestibility is the keyword here. Dietary fibers encompass a range of different compounds that affect many important gastrointestinal and systemic bodily processes (Brownlee, 2011). For example, the key role of the gut is to absorb nutrients following digestion in the stomach. Dietary fibers that inhibit this intestinal process results in decreased upper gastrointestinal transit times, in turn, affecting satiety. Furthermore, as complex carbohydrates, they take longer to digest because they require more time and effort to be broken down into their simpler sugar units. Dietary fiber is also known to decrease the risk of coronary heart disease (Abdul-Hamid and Luan, 2000) and other maladies (Peters et al., 2003). Additionally, indigestible seaweed polysaccharides like brown seaweeds (alginates) and edible red seaweeds (carrageenan's) are capable of forming ionic colloids that have been shown to lower serum⁵ cholesterol levels. This is in part due to their ability to disperse in water while retaining cholesterol and similar active compounds and to inhibit lipid absorption in the gastrointestinal tract (Jiménez-Escrig and Sánchez-Muniz, 2000).

    In short, fiber is a type of complex carbohydrate; it is a polysaccharide (made up of three or more sugar units) and is very difficult to digest. Fiber can be further classified as insoluble and soluble and is mainly composed of cellulose that can be found in whole grains, oats, peas, beans, root vegetables, potatoes, rice, and fruits including prunes, plums, and figs. Fiber is required by the body to promote digestive health.

    Cellulose—Physically, many plants rely on the cellulose polysaccharide for its structure and rigidity. As with starch, cellulose comprises beta-linked subunits or monomers⁶ of glucose. As such, they cannot be enzymatically broken down by humans and animals—that is, with the exception of certain ruminants (cows in particular) who utilize bacteria in the gut that can in fact hydrolyze the beta-link. For the rest of the animal kingdom though, cellulose usually passes through the digestive system largely undigested (Eliasson, 2016).

    Hemicelluloses—Hemicelluloses are composed of a mixture of pentose and hexose units (comprising five and six carbon atoms, respectively) including xylose, glucose, and mannose. Also, while they share certain similar characteristics to cellulose, including their names, they are in fact unrelated. Together with other complex carbohydrates like pectins, they surround the cellulose fibers of plant cells. Hemicellulose along with cellulose and pectin (a nonstarch polysaccharide), as well as lignin, a noncarbohydrate constituent of plants cell walls, collectively comprise the main components of dietary fiber as mentioned above. Also, as mentioned, dietary fibers are either water soluble or nonwater soluble. The water-soluble fibers include the gums, pectins, mucilages, and certain hemicelluloses. They are found in oats, barley, grapefruits, and the like and are generally good for the body's digestive processes. Most hemicelluloses along with cellulose and lignin, however, are not water soluble even though they can swell up with absorbed water. These substances are found in most leafy greens, and although cannot be digested, certain hemicelluloses can be partly assimilated by bacteria found along the gastrointestinal tract.

    Starches—Starches are without a doubt the principal carbohydrate in most diets, and like fibers, they are also complex carbohydrates and important constituents of the human diet. Chemically, starches are polysaccharides that consist of repeating glucose units. In origin, starches turn out to be the storage carbohydrate of many varied plants such as cereals and root vegetables. Starches are found in a partially granular crystalline form comprising two polymers⁷, namely, amylose⁸ and amylopectin⁹ (Eliasson, 2016). Some common cereal starches contain 15%–30% amylose starch, while others like corn, sorghum, rice, and barley contain largely amylopectin starch and are considered waxy. It is this amylose/amylopectin ratio within foods that is the key to the functional properties of starches. Starch crystalline structure is lost when it is heated in water (gelatinization¹⁰), as it allows for easier digestion to take place (Cummings and Stephen, 2007). What that means is when starch is dispersed in a liquid and heated, the liquid penetrates into the starch granule, swelling it until it is fully hydrated. Once hydrated (gelatinized), the swollen granules may increase the viscosity of the liquid and/or assist in forming gels. In very loose general terms, the amylose provides the gel strength, while the amylopectin provides the viscosity.

    Modified starch—Natural proportions of amylose and amylopectin in starchy foods are variable, and different cultivars of common species, like rice, display quite varied amylose-to-amylopectin ratios. However, genetic breeding can alter these ratios to influence properties like appearance and texture and viscosity or simply to improve gel stability and mouthfeel of the starch involved. Furthermore, chemical tampering can help reduce retrogradation, which is part of the process of staling of bread, lowering or increasing of gelatinization temperature, and enhancing freeze—thaw stability, controlling viscosity, improving acid and heat stability, and inhibiting gel formation. However, a note of caution is warranted; by altering the chemical nature of starch, the result can lead to it becoming resistant to digestion (Cummings and Stephen, 2007; Eliasson, 2016).

    So, in sum, it can be said that dietary carbohydrates are a diverse group of compounds with a range of chemical, physical, and physiological properties. While chiefly there for energy metabolism,¹¹ carbohydrates do indeed affect things like satiety, blood glucose and insulin, and lipid metabolism. In turn, carbs also have implications for overall health contributing to things like weight control, diabetes, and cardiovascular disease (Cummings and Stephen, 2007; Eliasson, 2016).

    1.1.1.4 Sugar alcohols

    A class of carbohydrates are the sugar alcohols or polyols. These are compounds that fall in between sugars and alcohols. While they are neither, they are in fact carbohydrates with structurally similar chemical compounds that partially resemble both sugar and alcohol. Common sugar alcohols include glycerine, sorbitol, mannitol, isomalt, xylitol, lactitol, erythritol, maltitol, maltitol syrup, and hydrogenated starch hydrolysates. Collectively, they are important in the food industry as sweeteners for all sorts of products. One of the main reasons for using polyols over traditional sugars is that they are generally fully metabolized by the body and subsequently contribute fewer calories than traditional sugars do. Calorie rates of the various sugar alcohols are in the range of 0.02–3 Cal/g as opposed to 4 Cal/g for traditional sugars. However, while most polyols not as sweet as sucrose, the exceptions, maltitol and xylitol, are about the same as sucrose.

    1.1.2 Proteins

    (See also Chapter 4)

    Proteinaceous compounds containing carbon, hydrogen, oxygen, and nitrogen and on occasion phosphorous and sulfur are among the most complex of the organic compounds found in nature. They are commonly found in plants and animals and form an essential part of a cells' protoplasm.¹² They are also a necessary part of the human diet too in which they help to build and repair various bodily tissues (Sawyer et al., 2002). Although proteins (polypeptides¹³) can readily supply energy if required to, their primary area of responsibility is that they are the key component of most cells, the muscle, connective tissues, hair, and the skin (Latham, 1997). In turn, every protein molecule is made up of numerous combinations of amino acids. The human body requires these amino acids to retain or replace existing proteins or to make new proteins. Any excess of ingested amino acids that aren't needed by the body are simply passed through urination. There are in fact two categories of proteins, those that the body can produce (nonessential proteins—amino acids) by itself and those that it cannot (essential amino acids), which in turn must be ingested. Of the hundreds of nonproteinogenic¹⁴ amino acids that exist, the body utilizes just 20 or so protein-making (proteinogenic) amino acids; eight of these, isoleucine, leucine, lysine, methionine, phenylalanine, threonine, tryptophan, and valine (plus a ninth, histidine for infants), cannot be synthesized by the body and are the essential amino acids that must be consumed as part of the diet (Latham, 1997; WHO, 1997; Beaton and Patwardhan, 1976; Wilson, 2008; Wu, 2013). The other amino acids including glycine, alanine, serine, cystine, tyrosine, aspartic acid, glutamic acid, proline, hydroxyproline, citrulline, and arginine must either be ingested as food or synthesized in the body (Wu, 2013). Proteins can be further classified as either complete or incomplete:

    Complete proteins—simply means a substance or food that contains all the essential amino acids a person requires. These include meat and animal products like milk and eggs (WFP, 2000). Although having said that while animal proteins offer complete essential amino acids, they are not necessarily the best source as animal proteins often contain unhealthy amounts of cholesterol and saturated fats (Wu, 2013).

    Incomplete proteins—are those foods that do not by themselves contain all the amino acids required by the body. They are generally of fruit and vegetable origin, and as such, foods are often combined to complete the full required compliment. However, there is one noteworthy exception among fruit and vegetables and that is the humble soybean. Soybean's amino acid profile is such that it is one of the very few vegetable sources considered a complete protein, supplying all of the essential amino acids needed (WFP, 2000; Tovar et al., 2002; Velasquez and Bhathena, 2007; Erdman, 2000; Wu, 2013).

    Proteins and satiety—General satiety can be explained by the fact that the stomach expands during eating, and as it does so, nerve receptors on the stomach wall sense the volume and pressure of the stomach contents. These feelings in turn convey signals to the brain, thus creating the sensation of fullness. Furthermore, vegetable-based proteins, as opposed to meat proteins, create the feelings of fullness; earlier, this means less vegetable protein needs to be ingested before one feels full (Kristensen et al., 2016). Satiety is a multifaceted construct. Studies cited by Chambers, McCrickerd, and colleagues (2016) in their paper on the satiating effects of foods suggest that cognitive and physiological sensory signals play important roles in determining whether one is and remains fully satiated. This hinges on responses by the body in terms of satiety signals that combine cognitive sensory perceptions with absorptive and ingestive signals that collectively determine the state of fullness. Indeed, classic research in this area promotes the notion of a hierarchy of satiating macronutrients starting with protein, going on to carbohydrate then fat (Chambers et al., 2016). This is not surprising as protein has undeniably been in the limelight for a long time now when it comes to investigative and real-life research about satiety. Certainly, supporting the idea of long-term weight loss through increasing one's protein intake, without altering overall energy consumption, cites Chambers et al. (2016), is a strategy that is supported by the literature. This is especially so if combined with the intake of dietary fiber, which has been shown to work though increasing the physical bulk of foods, an upsurge in gastric swelling, and slowing the flow of gastric voiding.

    The literature also supports a dietary combination of high protein and fiber foods containing more carbohydrate than fat (Chambers et al., 2016). It also, surprisingly, indicates that by manipulating the physical or textural qualities of foods and playing on the psychology of a food's sensory profile (claiming fullness and satiety), so one's perception of satiety is also altered in favor of being full (Chambers et al., 2016).

    1.1.3 Fats

    (See also Chapter 16)

    Semantics—While generally referred to as fats, these substances are in fact lipids. They are fatlike compounds that encompass naturally occurring molecules that include fats, waxes, sterols (including cholesterol), fat-soluble vitamins (A, D, E and K), and phospholipids¹⁵.

    Fats and oils, collectively known as triglycerides¹⁶ or triglycerols, are commonly referred to in this sense as dietary fats. They are obtained from both animal and vegetable origin. Lipids make up the bulk of the dietary fat intake. In common with general usage, the term fat or fats in the context of this section is used to represent all edible fats and oils (triglycerides), while the term dietary fats (distinct from dietary lipids) is also used interchangeably with fat or fats—once again in line with common parlance (Latham, 1997; WFP, 2000; WKU, 2010). When it comes to fat, the difficulty for us humans is that, while we ingest and store fats, fats can also be metabolized from excess carbohydrate intake as well. Thus, one has to be careful and watch both fat and carbohydrate intake. Most dietary fats are composed of some sort of fatty acid¹⁷ arrangement; triglycerides are no different in this respect and are made of glycerol (a type of alcohol) and three fatty acids (Gibson, 2012). This separation is important as the nature of the fat depends on the specific type of fatty acids present. These can be divided into one of two groups: saturated fatty acids and unsaturated fatty acids, abbreviated to saturated and unsaturated fats, respectively. While it is true to say that all fats contain both saturated and unsaturated fatty acids, they are usually identified according to the dominant fatty acid present (Latham, 1997; WKU, 2010; Ortega, 2007).

    Saturated fats—Broadly speaking, these dietary fats derive predominantly from animals and animal by-products, for example, meat, eggs, milk, cream, and butter. Although, as always, there are exceptions particularly from the vegetable world; these include palm and coconut. Saturated fats are also differentiated from their unsaturated counterparts in that saturated fats have no carbon double bonds and are instead saturated with hydrogen atoms. A cautionary note, moderation must be exercised here as diets containing too much saturated fat tend to raise blood cholesterol levels affecting potential heart disease.

    Unsaturated fats—Unsaturated fats predominantly originate from plants and are usually further subdivided into two types: mono- and polyunsaturated fats. While both help in lowering low-density lipoproteins (LDL cholesterol) and reducing heart disease, some consider polyunsaturates as doing the job marginally better by providing more membrane fluidity than monounsaturated fats. Unsaturated fats in general comprise carbon-to-carbon double bonds, the number and configuration of which ultimately determines health properties in the humans. Monounsaturated fats contain only one such bond, while polyunsaturated fats contain more than one. The healthy unsaturated fatty acids are those with the cis as opposed to the trans carbon configuration (WKU, 2010; WHO/EMRO, 2010; EFSA, 2008). Both the following are of the cis configuration:

    Monounsaturated fats (MUFA) include oils from peanuts, olives, canola, sesame seed, and avocados. It's worthwhile mentioning too that monounsaturated fats are also high in vitamin E, which also helps to develop and maintain healthy cells in the body.

    Polyunsaturated fats (PUFA) include oils from corn, cottonseed, safflower, soybean, and sunflower. Fish oil too is also polyunsaturated as are omega-3 and omega-6 fatty acids.

    Although for some, polyunsaturated fats are the preferred choice; they are more vulnerable to peroxidation (rancidity) (Section 16.1). Although in an attempt to reduce and slow the onset of rancidity and prolong fats' shelf life, scientists introduced hydrogenation.¹⁸ Unfortunately, however, partial hydrogenation, while beneficial in the reduction of rancidity and allowing fats and oils to be firm/hard at room temperature, hydrogenizing polyunsaturated fats in part transforms the original oil/fat into trans-fatty acids or trans fats for short (WKU, 2010; WHO/EMRO, 2010; EFSA, 2008).

    Trans fats—Trans fats can be either trans-PUFA or trans-MUFA, and while they both can occur naturally in nature, they are not common. Instead, by far, the biggest producers and users of trans fats are humans. Trans fats are often found in processed foods such as soft or semihard margarines. They are also particularly harmful in that trans fats act like saturated fats (see note on hydrogenation) in that they can raise levels of cholesterol in the blood and just as disturbing contributes to fatty acid membrane loss in the brain with implications for neurodegenerative disorders such as multiple sclerosis (MS) and Parkinson's and Alzheimer's diseases (WKU, 2010; WHO/EMRO, 2010; EFSA, 2008).

    Essential fatty acids—When required, the human body can produce (synthesize) most of the fatty acids it needs; in this respect, they are not essential fatty acids. Others, however, the essential fatty acids, are essential and must be included in the diet. Two essential fatty acids, linoleic and alpha-linolenic, cannot be manufactured in the body and therefore must be obtained from consumed foods such as vegetables and some vegetable oils (flaxseed), nuts (walnuts), seeds (pumpkin and sunflower), soya and soya products, and fish and fish oils. Linoleic and alpha-linolenic are used to build particular fats—omega-3 and omega-6 fatty acids that aid in the normal functioning of all tissues of the body (Gropper and Smith, 2012). However, it is also important to, not only ingest these fatty acids, but just as importantly to take them in the right ratios. This is because both fatty acids are competitive in nature and therefore, excessive intake in one might inhibit the effect of the other (Lands et al., 1990). The ideal ratio of omega-6 to omega-3 should be in the region of 1:1 and 4:1 (Simopoulos, 1999). Instead, many consume these fats at a much greater ratios—as high as 10:1 and 25:1. Such a runaway imbalance is due in large part because of the greater dependence on processed foods and oils. Aiming for a lower omega-6/omega-3 ratio is desirable for reducing the risk of many chronic diseases (Simopoulos, 2008).

    1.2 Water

    Water is often not seen as or is more precisely overlooked as a macronutrient, yet it is perhaps one of the most important of the macronutrients. While water does not provide the body with energy like fats, proteins, and carbohydrates do, it does however account for around 60–70% of our total body weight (Fig. 1.1) (EUFIC, 1998; DeMan, 1976). And we cannot survive without it. Water plays an important role in many nutritional processes, not least of which is homeostasis.¹⁹ Water and homeostasis work in unison regulating the bodies fluids; it controls the body's temperature (via perspiration), aids in digestion, and enables the transportation of nutrients around the body; water is also involved on a chemical level too, to facilitate the metabolization of foods, and lastly, it also promotes efficient bodily waste management through fecal and urine excretion (WHO/EMRO, 2010; CFNI, 2004; Lipp et al., 1999; Matthys et al., 2007; Wilson, 2008; EUFIC, 2009).

    Fig. 1.1 Nutrient composition of the human body by weight.

    Losing just 8% of the average person's body water (through dehydration, heatstroke, constant diarrhea, etc.) can have dire consequences; losing 10% can be fatal (EFSA, 2008). Yet for all its importance, water intake requirements were still, up until recently at least, prescribed solely on the need to prevent such things as dehydration and little else (Grandjean, 2004). After more research in the area, however, it has been calculated that just about 25–35 mL of fluids per kilogram body weight must be consumed daily—which is equivalent to between 2 and 3 L/day for an adult (Howard and Bartram, 2003). In short, without food, humans can survive many weeks by calling on its reserves, yet without water, humans can only survive for just a few short days (CFNI, 2004; Belitz et al., 2009). The main sources of water are from what we drink plus that found in foods we ingest.

    1.3 Micronutrients

    Unlike macronutrients, micronutrients provide virtually no energy to the body and as such are only required in small does. While the term micro-nutrients might be suggestive of lesser importance, it must be remembered that micro is a volumetric term only; consequently, micronutrients should not be used to relegate to a position of secondary value. Essential micronutrients are involved in many processes in the body, without which, even in these small amounts, would result in severe problems in the body. Micronutrients promote biochemical reactions like enzyme production and aid in metabolization and producing hormones and many other compounds or substances, all of which are essential for the proper regulation of growth and development within the human body (EUFIC, 2009; WHO, 1997, 2017). Micronutrients consist primarily of vitamins, minerals, and trace elements.

    1.3.1 Vitamins

    Vitamins are organic substances that are produced by plants and animals; they fall neatly into two groups: fat or water soluble. Fat-soluble vitamins including A, D, E, and K can be stored in the body, so they need not be ingested so frequently. Water-soluble vitamins on the other hand, like vitamin C and the B complex varieties (vitamins B1 (thiamine), B2 (riboflavin), B3 (niacin), B6, B9 (folic acid or folate), and B12), for example, need to dissolve in water before the body can absorb them. Because of this solubility, the body cannot store them as reserves, as a result any excesses are passed through the urinary tract. Consequently, the body requires a fresh supply of many of these micronutrients every day (Latham, 1997; WHO, 1997).

    Although many need replenishing every day, vitamins are only required in very small amounts. They also have diverse biochemical functions. Some have hormonelike functions as regulators of mineral metabolism (vitamin D), or regulators of cell and tissue growth and differentiation (some forms of vitamin A). Others function as antioxidants (vitamin E and sometimes vitamin C). While the largest number of vitamins (the B complex vitamins) function mainly as precursors for enzyme cofactors (coenzymes²⁰) that aid enzymes in their work as catalysts in metabolism. Other functions of vitamins include boosting the immune system while helping cells and organs do their jobs. Vitamin K helps blood to clot (Yang and Xiong, 2012). Furthermore, if one is running on empty and showing signs of fatigue, tiredness, and low energy levels, despite sufficient macronutrient intake and without any underlying disease, such symptoms might just as easily be caused by a lack of vitamins and minerals in the diet (see also Appendix A and Section 1.1) (Huskisson et al., 2007).

    1.3.2 Minerals

    As opposed to vitamins and proteins, minerals are inorganic chemical elements that are needed by all living beings. Yet as with vitamins, minerals are often relegated to a position of secondary status. However, this cannot be further from the truth as they provide important functions either by themselves or in conjunction with other micro- and macronutrients. Minerals are indirectly provided through plants and animals, which in turn ingest them via soil and water or, in the case of animals, through eating plants or other animals. Minerals are also carried in the water supply, the composition of which varies greatly from region to region (LPI, 2017). Moreover, it is also worth noting that certain elements including some metals occur in the form of ionic²¹ compounds. Minerals are acquired, not only necessarily individually but also in complex compounds via natural inorganic sources. In other cases, certain minerals occur less frequently in nature and have to be added artificially in the form of supplements like iodine in salt, for example. There are some minerals that are essential, and they have to be consumed in larger amounts. Sometimes, these minerals are referred to as macrominerals; other times, this distinction is not made. Mineral's roles can be structural as in bones and teeth, for example, or perhaps in vital roles such as electrolytes²² or more simply to help regulate many of the body's processes including metabolization and energy production; others might have more specific functions like providing structure in (FSA, 2010; MIT, 2010; WHO/EMRO, 2010; WHO/FAO, 2004).

    As with many things, mineral data vary from individual to individual and from institution to institution. Bearing this in mind, minerals can be categorized either by their importance or functional classification, by status of essential and nonelements, by their structural compound such as metals and salts, or simply by groupings like macro- and micronutrients and macro- and microminerals and trace minerals (MIT, 2010; WHO/FAO, 2002).

    This book follows the latter convention. However, as has been pointed out, just as there is a lack of agreement between people and institutions on macro and micronutrients, so there is a similar lack of mutual agreement between the separation of macro- and microminerals. That said, some system must be adopted, so this author chooses to recognize macromineral nutrients, as those needed in quantities above 100 mg and microminerals as those between 1 and 100 mg. Trace elements too, for the purposes of this book, are those that fall below this 1 mg threshold (Britannica, 2009; MIT, 2010). Importantly, while certain minerals are required in larger doses than others, this does not confer upon them a greater measure of importance. Regardless of which system is used, we still need some perspective, guidelines, and recommendations on daily intake. This is where the dietary guidelines come in (next section and Appendix A) (FAO/WHO, 2001; Bogden and Klevay, 2000).

    1.3.3 A Good Balanced Energy Portfolio

    When it comes to energy intake, it is not sufficient to simply ingest sufficient calories to meet daily requirements. Instead, because of the body's changeable physiological needs and the preferential way it treats the different energy macronutrients (carbohydrates, fats, and proteins), a balanced energy portfolio is recommended by all major dietary guideline providers. Deciding adequate protein, carbohydrate, and fat balances is a complicated and costly exercise. Having said that, it is nevertheless important as imbalances can inevitable be costly in terms of health both from a personal and from a public health perspective (WHO/FAO/UNU, 2007; WFP, 2009). It can be seen from the appendices (Section A.4 and Table A.4) that carbohydrates, the major source of energy, can be as high as 75% of total macronutrient intake. It can also be seen from Section A.4 and Table A.3 that the figures indicate that both industrialized and nonindustrialized countries' daily fat intake can be as high as 35% (Latham, 1997). This can particularly skew energy intake as is seen in the discussion on metabolism (Section 1.4). It is also worth noting at this juncture that the energy content of fat equals 9 kcal/g, which is more than twice that of protein and carbohydrate (both at 4 kcal/g). Therefore, any intake above and beyond expenditure, as quoted by national guidelines, will result in this fat being deposited as stored fat (usually as adipose tissue). So, what are the recommended balances and who decides this? For a fuller answer, see Appendix A Dietary Guidelines. In the meantime, we take a quick look at the body's metabolic processes vis-à-vis the macronutrients.

    1.4 Metabolism

    Unlike plants (autotrophs²³) that are able to convert sunlight into energy via photosynthesis, so animals including humans (heterotrophs²⁴) rely on organic molecules or compounds (food) as starting blocks for energy and other purposes. As mentioned, foods contain six major nutrient classes: carbohydrates, proteins, fats, vitamins, minerals, and water. Between them, they provide the raw materials the body requires for energy and for building cells like proteins, certain carbohydrates, lipids, nucleic acids such as DNA and RNA, and other important components of the body. These breaking down and building up of biochemical compounds are referred to as metabolism or sometimes cellular respiration²⁵. Put simply, metabolism is the sequence of chemical reactions that happens within a cell. It is the biochemical breakdown of complex molecules in a process called catabolism resulting in the breakdown of larger molecules into smaller ones while releasing energy in the process or the utilization of the energy of catabolized molecules for building or synthesizing other molecules in a sequence of metabolic pathways called anabolism. The various ways or paths that are available to achieve the entire process of metabolism are referred to as the metabolic pathways (more later).

    However, before metabolism proper can take place, the food we eat must first be digested; this is sometimes called intermediary or intermediate metabolism. Digestion and metabolism also involve the movement or transportation of substances and the like between and into the many varied cells (Gropper and Smith, 2012). It all begins with the digestive process, in particular, with the digestive enzymes that first break down foods into their component building blocks allowing for easier absorption into the body. The body achieves this, in part, by utilizing enzymes, proteinaceous molecules that catalyze or speed up certain chemical processes of metabolism. Specific enzymes work in specific environments, temperatures, and pH levels. Digestive enzymes are located all along the digestive tract, in the salivary glands, in cells that line the stomach, in pancreatic secretions, and from cells in both the large and small intestines. Nutrients are also broken down in cells at the cellular level. Enzymes tend to be categorized according to their particular area of speciality. Therefore, we can say that the enzyme amylase,²⁶ which is produced in the salivary glands, the pancreas, and the small intestine, aids in the breakdown of carbohydrates like starch and sugars into simple sugars like glucose—both in the mouth and in the small intestine. All the while, protease enzymes catalyze proteins into their constituent components (i.e., peptides and amino acids) in the stomach and small intestine. Lipases on the other hand are found in the pancreas and small intestine and accelerate the hydrolysis of lipids resulting in fatty acids and glycerol in the small intestine. Lastly, nucleases split the genetic transporters, the nucleic acids into nucleotides. Taking this one step further, foods are primarily broken down (catabolized) into food molecules (nutrients) in the digestive tract. It does this by initially using the mechanical forces of biting and chewing of food in the mouth; at the same time, the salivary glands secrete amylase. This helps digest or break down starch before the food enters the stomach where hydrochloric acid and pepsin²⁷ further break down various nutrients. This mix of food, saliva, and gastric juices is collectively known as chyme. This chyme then travels to the small intestine where the pancreas produces proteolytic lipases and a plethora of other digestive enzymes, which, together with the liver's bile salts, further digest (break down) nutrients for easier ingestion. After or, more precisely, during the small intestinal pathway, glucose, simple carbohydrates, and other nutrients are carried across the intestinal wall and absorbed directly into the blood. At this point, the liver's parenchymal cells along with other tissues help to finally complete the initial catabolic process by breaking down the remaining carbohydrates into simple sugars, fats into fatty acids, and proteins into amino acids. After this, nutrients are then made available to the cells at which point, if need be, these nutrients can be further altered or broken down into numerous other chemicals for the cell's requirements (King, 2010).

    Moving on, howsoever, the food/nutrients are processed or digested; the metabolites (as the products of metabolism are called) are transported via the bloodstream (produced in the bone marrow) to where they need to be. The bloodstream comprises two components, cells and plasma. The cellular portion of blood contains three categories of cells: red blood cells (RBCs) (also known as erythrocytes), whose purpose is to deliver oxygen to the body's cells while taking away spent CO2. Another blood cell component is the white blood cells (WBCs), which aids in the fight against infection. Then, there are the platelets (also known as thrombocytes), which are small disk-shaped colorless cell fragments without a nucleus found in large numbers in blood and which are involved in the clotting process. Lastly, the plasma (the clear yellowish blood fluid when the cells are separated out) is responsible, not only as a vehicle for carrying the cells but also, just as importantly, as a transport carrier for such things as dissolved electrolytes, nutrients, vitamins, hormones, clotting factors, and antibody²⁸ proteins such as albumin and immunoglobulins (Belitz et al., 2009; Gropper and Smith, 2012).

    When food molecules are not being metabolized for specific purposes in the body, they are metabolized for energy. These metabolic reactions are the chemical reactions in which living organisms convert chemical compounds (metabolized food) into useable energy or other components needed by the body. Of the numerous pathways involved in metabolizing food molecules, there are a handful of important metabolic pathways worth noting; these are shown in Table 1.2.

    Table 1.2

    Source: Compiled from multiple sauces: Gropper, S.S., Smith, J.L., 2012. Advanced Nutrition and Human Metabolism. Cengage Learning; Devlin, T.M., 2010. Textbook of Biochemistry With Clinical Correlations. John Wiley & Sons, USA; Greenberg, D., 2012. Metabolic Pathways. Elsevier; Medeiros, D.M., Wildman, R.E.C., 2013. Advanced Human Nutrition. Jones & Bartlett Learning.

    1.5 Energy Preferences

    The body's first requirement is for energy and, if required, all macronutrients, that is, proteins, fats, and carbohydrates, with the exception of water, to ultimately be catabolized into functional glucose, glycerol, and free fatty acids (fats). That said, this interchangeability of energy sources is not as might seem, a randomized process; instead, the human body has its preferences. With its relatively quick and easy transformation into glucose, carbohydrates are the major source of energy for the body. In this sense, carbohydrate (in the form of glucose) is the body's fuel of choice. In the absence of sufficient intake, or a bout of excessive physical activity whereby reserves of glucose run low, then the body first taps into its glycogen reserves. Once these are depleted or become dangerously low, so the next place the body turns to are the fatty deposits (the adipose cells). If still in need of energy, through prolonged abstinence, profoundly low continuous intake or starvation, then the body will start to convert its own existing muscle proteins into its constituent amino acids and then into glucose for energy. This has the effect of depleting muscle mass, and action must be taken to remedy this.

    Just as glucose is the body's choice of fuel, so adipose (or fat) tissue is the body's choice of fuel storage system; this is no accident. As with protein, adipose is very flexible and either can be synthesized from excess dietary fat intake or can be converted from surplus carbohydrate or protein in our body and/or diet. As it does so, so fat reserves are built up in case of leaner times. The first and foremost reason the body converts all this extra carbohydrate, protein, and fat into body fat is the relative energy profile of fat itself. That is to say, each type of molecule, whether fats, carbohydrates, and proteins, produces specific amounts of energy. For example, the energy yield from 1 g of fatty acid is approximately 9 kcal (37 kJ), compared with carbohydrates and proteins, which is just 4 kcal/g (kJ/g) each. Thus, it is for this reason that fats are so potent and efficient forms of energy stores. Hence, fat stores can be converted back into glucose as and when needed (Wilson, 2008).

    1.5.1 Carbohydrate Metabolism

    As mentioned, carbohydrates are the body's first choice when it comes to energy production within the body (WFP, 2000). Carbohydrate metabolism utilizes several of the abovementioned pathways in the production of glucose (also referred to as blood sugar), the cells' universal source of energy. Essentially, the human body, through the digestive tract or GI, tends to treat all carbohydrates in the same way. That is, it endeavors to break down carbohydrates into sugar molecules whereupon it converts the most digestible of these into glucose (Fig. 1.2) (HSPH, 2010). In the carbohydrate metabolic pathway, the glucose is absorbed into the cell directly from the blood. From here, it enters the cytosol (the aqueous liquid inside the cell) where through the initial process of glycolysis (the oxidation of glucose) converts glucose molecules into two molecules of pyruvic acid (pyruvate) and a modest amount of the universal energy storage molecule—adenosine triphosphate (ATP). The pyruvate then passes into the mitochondria²⁹ from the cytosol (Alberts et al., 2002). Two options are present at this juncture dependent on the cell's needs; this in turn determines whether the pyruvate is subsequently treated aerobically or anaerobically. In anaerobic activity, the pyruvic acid is subject to anaerobic glycolysis³⁰ (Hunter, 2008). If sufficient oxygen is present for aerobic conversion, pyruvic acid is further metabolically transformed into acetyl-CoA before it enters the citric acid cycle (also known as the Krebs cycle). In the meantime, oxygen from the lungs is dispersed into the blood. Oxygenated blood is then fed to the cell's mitochondrion. In the citric acid (the Krebs) cycle, the oxygen in the blood is made available to the cell and in turn reacts with the acetyl-CoA oxidizing it and generating a small amount of energy that is then stored as nicotinamide adenine dinucleotide hydrogen (NADH) and waste carbon dioxide (Hunter, 2008). The spent carbon dioxide travels back, via the blood, to the lungs where it is exhaled, while the hydrogen continues on in other reactions to further form adenosine triphosphate (ATP) molecules and water (Fig. 1.2).

    Fig. 1.2 Foods' metabolic pathways.

    The ATP molecule acts like a miniscule battery allowing it to be used for either molecular growth or energy release. If the cell requires molecular growth or repair (called biosynthesis or biogenesis), the ATP molecule anabolizes or recombines simple molecules into complex ones; consequently, one has cellular growth or repair. If on the other hand, energy is required for movement or the metabolic process itself, the ATP molecule is degraded back to adenosine diphosphate (ADP), and sufficient energy is released for the required purpose. When the organism or cell is at rest and energy is actually not immediately required, in this case, the reverse reaction takes place, thus storing energy like a battery (Fig. 1.2).

    Any excess intake of carbohydrate is first converted to glucose then stored as glycogen—a complex carbohydrate in skeletal muscle and the liver or as fat (triglycerides in adipose tissue) providing a reserve storage of energy. As and when energy stores are again needed, so glycogen is the first reserve to be called upon, and then, if further energy is required and glycogen is depleted, the body then reuses the triglycerides (fat) (Albright and Stern, 1998; UoArizona, 2004; Britannica, 2009; King, 2010; Silva, 2002; UoAkron, 2010; UoBristol, 2010).

    1.5.2 Lipid (Fat) Metabolism

    As well as carbohydrates, the body requires other macronutrients like lipids (fats) and protein for energy and growth. And as with carbohydrates, any intake in excess of the body's needs is converted and stored in the body as triglycerols (or triglycerides). Then, in response to cellular energy demands, ingested fatty acids or stored fats can be utilized. Initially triglycerols are catabolized (broken down) in an enzymatic process of lipolysis into fats' component parts, that is, free fatty acid and glycerol. The free fatty acid and glycerol are treated separately. Free fatty acid is oxidized in the mitochondrion organelles³¹ of liver cells in a process called β-oxidation.³² This leads to the formation of NADH and the energy battery ATP. The glycerol on the other hand is converted to glucose in a process of gluconeogenesis,³³ which then becomes available to the citric acid cycle and subsequently is metabolized into NADH and ATP (Merck Manual, 2005; King, 2010; Albright and Stern, 1998; DeMan, 1976).

    1.5.3 Protein Metabolism

    Each protein comprises as many as 2000 amino acids connected by chemical links called peptide bonds. So, for this large molecule to be used in the body, it must first be catabolized into its component amino acids. This might sound a little inefficient, as amino acids usually also act as inputs for the building of the body's own proteins. Protein turnover throughout the body, for instance, including new proteins being synthesized and existing proteins being broken down, takes place constantly. However, apparent inefficiency aside, if needed for energy, amino acids are first converted to glucose by gluconeogenesis, primarily in liver cells. These are then converted into pyruvate through glycolysis before entering the citric acid cycle for oxidation and ultimately, the production of ATP. If intake, however, is in excess of the body's natural requirements, proteins are then metabolized to glycogen or triglycerols (fat) and are subsequently stored for use later as needed (Latham, 1997).

    References

    Abdul-Hamid A., Luan Y.S. Functional properties of dietary fibre prepared from defatted rice bran. Food Chem. 2000;68(1):15–19.

    Alberts B., et al. Molecular Biology of the Cell: International Student Edition. San Francisco: Routledge; 2002.

    Albright A.L., Stern J.S. Encyclopedia of Sports Medicine and Science. Will Hopkins; 1998. http://www.sportsci.org/encyc/adipose/adipose.html.

    Beaton G.H., Patwardhan V.H. Physiological and practical considerations of nutrient function and requirement. In: Beaton G.H., Bengoa J.M., eds. Nutrition in Preventative Medicine: The Major Deficiency Syndromes, Epidemiology and Approaches to Control. Geneva: World Health organisation; 1976.

    Belitz H., et al. Food Chemistry. 4th ed. revised and extended edition New York, Philadelphia: Springer; 2009.

    Bogden J.D., Klevay L.M. Clinical Nutrition of the Essential Trace Elements and Minerals: The Guide for Health Professionals. Totowa, NJ: Humana Press; 2000.

    Britannica, 2009. Encyclopaedia Britannica Online.

    Brownlee I.A. The physiological roles of dietary fibre. Food Hydrocoll. 2011;25(2):238–250.

    CFNI. Water the Forgotten but Essential nutrient. Jamaica: Caribbean Food and Nutrition Institute; 2004.

    Chambers L., et al. Optimising foods for satiety. Trends Food Sci. Technol. 2016;41(2):149–160.

    Cummings J., Stephen A. Carbohydrate terminology and classification. Eur. J. Clin. Nutr. 2007;61:S5–S18.

    DeMan J.M. Principles of Food Chemistry. Westport, CN: AVI Publishing Co. Inc., 1976.

    EFSA. Draft: dietary reference values for water. Scientific Opinion of the Panel on Dietetic Products, Nutrition and Allergies. Parma: European Food Safety Authority; 2008.

    Eliasson A.-C. Carbohydrates in Food. Boca Raton, FL: CRC Press Taylor & Francis Group; 2016.

    Erdman J.W. AHA science advisory: soy protein and cardiovascular disease—a statement for healthcare professionals from the nutrition committee of the AHA. Circulation. 2000;102(20):2555–2559.

    EUFIC. The European Food Information Centre: What Do We Mean by Nutrition? Retrieved 11 November 2016 from http://www.eufic.org/article/en/health-and-lifestyle/healthy-eating/artid/nutrition-2/. 1998.

    EUFIC. The European food information centre: the basics: nutrition. Retrieved 15 November 2016 from http://www.eufic.org/article/en/page/BARCHIVE/expid/basics-nutrition/. 2009.

    FAO/WHO. Bankok, Thailand: FAO/WHO; . Report of the Joint FAO/WHO Expert Consultation on Human Vitamin and Mineral Requirements. 2001;303.

    Field S.Q. Culinary Reactions: The Everyday Chemistry of Cooking. Chicago: Chicago Review Press; 2011.

    FSA. Healthy Diet. Available from: http://www.eatwell.gov.uk/healthydiet/. 2010.

    Gibson M. The Feeding of Nations: Re-defining Food Security for the 21st Century. Boca Raton, FL: CRC Press; 2012.

    Grandjean A. Water Requirements: Impinging Factors and Recommended Intakes. Geneva: World Health Organisation; 2004.

    Gropper S.S., Smith J.L. Advanced Nutrition and Human Metabolism. Boston: Cengage Learning; 2012.

    Howard G., Bartram J. Domestic Water Quantity, Service Level and Health. Geneva: WHO/SDE/WSH; . 2003;39.

    HSPH. The nutrition source. Retrieved 5 October 2016 from http://www.hsph.harvard.edu/. 2010.

    Hunter A.D. Chemistry 1506: Allied Health Chemistry 2 Section 12: Specific Catabolic Pathways, Molecular Destruction. Youngstown, OH: Department of Chemistry, Youngstown State University; 2008.

    Huskisson E., et al. The role of vitamins and minerals in energy metabolism and well-being. J. Int. Med. Res. 2007;35(3):277–289.

    Jiménez-Escrig A., Sánchez-Muniz F.J. Dietary fibre from edible seaweeds: chemical structure, physicochemical properties and effects on cholesterol metabolism. Nutr. Res. 2000;20(4):585–598.

    King M.W. The Medical Biochemistry Page. Retrieved 6 November 2016 from http://themedicalbiochemistrypage.org/home.html. 2010.

    Kristensen M.D., et al. Meals based on vegetable protein sources (beans and peas) are more satiating than meals based on animal protein sources (veal and pork)–a randomized cross-over meal test study. Food Nutr. Res. 2016;60:32634.

    Lands W.E., et al. Quantitative effects of dietary polyunsaturated fats on the composition of fatty acids in rat tissues. Lipids. 1990;25(9):505–516.

    Latham M.C. Human Nutrition in the Developing World. Rome: Food and Agriculture Organisation of the United Nations; 1997. Food and Nutrition Series—No. 29.

    Lee S.-Y., et al. Factors influencing the stability of garlic thiosulfinates. Food Sci. Biotechnol. 2014;23(5):1593–1600.

    Lipp J., et al. Techniques and procedures; fluid management in enteral nutrition. Nutr. Clin. Pract. 1999;14(5):232–237.

    LPI. Micronutrient information center. Retrieved 12th April 2017 from http://lpi.oregonstate.edu/. 2017.

    Matthys C., et al. Beverage consumption in Belgian adolescents. Asia Pac. J. Clin. Nutr. 2007;16(suppl. 3):S58.

    Merck Manual. In: Porter R.S., Kaplan J.L., eds. Inherited Disorders of Metabolism: Fatty Acid and Glycerol Metabolism Disorders. New Jersey: Merck Sharp & Dohme Corporation; 2005 (Online Medical Library).

    MIT. Optimizing Your Diet. Retrieved 30 October 2010, from http://web.mit.edu/athletics/sportsmedicine/wcrminerals.html. 2010.

    Ortega J.B. Polyunsaturated Fatty Acid Metabolism in Broiler Chickens: Effects of Maternal Diet. (Master of Science) Oregon: Oregon State University; 2007.

    Peters U., et al. Dietary fibre and colorectal adenoma in a colorectal cancer early detection programme. Lancet. 2003;361(9368):1491–1495.

    Sawyer C.N., et al. Chemistry for Environmental Engineering and Science. Avenues of the America's, NY: McGraw-Hill Education; 2002.

    Silva P. A General Overview of the Major Metabolic Pathways. Retrieved 21 November 2010 from http://www2.ufp.pt/~pedros/bq/integration.htm. 2002.

    Simopoulos A.P. Essential fatty acids in health and chronic disease. Am. J. Clin. Nutr. 1999;70(3):560s–569s.

    Simopoulos A.P. The importance of the omega-6/omega-3 fatty acid ratio in cardiovascular disease and other chronic diseases. Exp. Biol. Med. 2008;233(6):674–688.

    Tovar A.R., et al. A Soy protein diet alters hepatic lipid metabolism gene expression and reduces serum lipids and renal fibrogenic cytokines in rats with chronic nephrotic syndrome. J. Nutr. 2002;132:2562–2569.

    UoAkron. Department of Chemistry: Carbohydrate Metabolism. Akron, OH: Hardy Research Group: The University of Akron; 2010.

    UoArizona. University of Arizona: The Biology Project. Arizona, OH: The University of Arizona; 2004.

    UoBristol. Acetyl Coenzyme A: The Molecule That Makes Fats, or Burns Them by Paul May. Bristol: University of Bristol; 2010.

    Velasquez M.T., Bhathena S.J. Role of dietary Soy protein in obesity. Int. J. Med. Sci. 2007;4(2):72–82.

    WFP. Food and Nutrition Handbook. Rome: World Food Programme (WFP); 2000.

    WFP. Emergency Food Security Assessment Handbook. Rome: World Food Program; 2009.

    WHO. Nursing Care of the Sick: A Guide for Nurses Working in Small Rural Hospitals. Manila, Philippines: World Health Organization; 1997. Western Pacific Education in Action Series No. 12.

    WHO. Neurological Disorders: Public Health Challenges. Geneva: World Health Organization; 2006.

    WHO. Website of the World Health Organisation. Retrieved 2 April 2017, from http://www.who.int/nutrition/topics/5_population_nutrient/en/. 2017.

    WHO/EMRO. You Are What You Eat. World Health Organisation: Eastern Mediterranean Regional Office; 2010.

    WHO/FAO. Vitamin and Mineral Requirements. Rome: WHO/FAO; 2002.

    WHO/FAO. Vitamin and Mineral Requirements in Human Nutrition (Draft Version). Sun Fung, China: WHO/FAO; 2004.

    WHO/FAO/UNU. Joint FAO/WHO/UNU Expert Consultation on Protein and Amino Acid Requirements in Human Nutrition. Geneva: World Health Organisation; 2007. WHO Technical Report Series 935.

    Wilson M.M.G. Disorders of nutrition and metabolism. In: Porter R.S., Kaplan J.L., eds. The Merck Manual Online Medical Library. New Jersey: Merck Sharp & Dohme Corporation; 2008.

    WKU. Bio 113- Lipids. Retrieved 15 July 2010 from http://bioweb.wku.edu/courses/biol115/Wyatt/Biochem/Lipid/lipid1.htm. 2010.

    Wu G. Amino Acids: Biochemistry and Nutrition. Boca Raton, FL: CRC Press; 2013.

    Yang Z., Xiong H.-R. Culture Conditions and Types of Growth Media for Mammalian Cells. In: Ceccherini-Nelli L., ed. Biomedical Tissue Culture. InTech; 2012:doi:10.5772/52301. Available from: https://www.intechopen.com/books/biomedical-tissue-culture/culture-conditions-and-types-of-growth-media-for-mammalian-cells.

    Further Reading

    Devlin T.M. Textbook of Biochemistry With Clinical Correlations. New York: John Wiley & Sons; 2010.

    Eduweb. Fact Sheets: Body Composition. Retrieved 15 October 2017 http://www.rowett.ac.uk/edu_web/. 2010.

    Gibney M.J., et al. Introduction to Human Nutrition. Oxford: Blackwell Publishing; 2009.

    Greenberg D.M. Metabolic Pathways: of Chemical Pathways of Metabolism. Salt Lake City: Academic Press; 2012.

    Medeiros D.M., Wildman R.E.C. Advanced Human Nutrition. Burlington, MA: Jones & Bartlett Learning; 2013.


    ¹ Fructose is the sweetest of all the sugar carbohydrates.

    ² Hydrolysis is the breakdown of a compound by a chemical reaction with

    Enjoying the preview?
    Page 1 of 1