Nov 25, 2008

CARBOHYDRATE CHEMISTRY

The function of carbohydrates includes energy storage and providing structure. Sugars are carbohydrates, although there are carbohydrates that are not sugars. There are more carbohydrates on Earth than any other type of biomolecule. The simplest type of carbohydrate is a monosaccharide, which among other properties contains carbon, hydrogen, and oxygen, mostly in a ratio of 1:2:1 (generalized formula CnH2nOn, where n is at least 3). Glucose, one of the most important carbohydrates, is an example of a monosaccharide. So is fructose, the sugar that gives fruits their sweet taste. Some carbohydrates (especially after condensation to oligo- and polysaccharides) contain less carbon relative to H and O, which still are present in 2:1 (H:O) ratio. Monosaccharides can be grouped into aldoses (having an aldehyde group at the end of the chain, e. g. glucose) and ketoses (having a keto group in their chain; e. g. fructose). Both aldoses and ketoses occur in an equilibrium between the open-chain forms and (starting with chain lengths of C4) cyclic forms. These are generated by bond formation between one of the hydroxy groups of the sugar chain with the carbon of the aldehyde or keto group in a semiacetal bond. This leads to saturated five-membered (in furanoses) or six-membered (in pyranoses) heterocyclic rings containing one O as heteroatom. Two monosaccharides can be joined together using dehydration synthesis, in which a hydrogen atom is removed from the end of one molecule and a hydroxyl group (—OH) is removed from the other; the remaining residues are then attached at the sites from which the atoms were removed. The H—OH or H2O is then released as a molecule of water, hence the term dehydration. The new molecule, consisting of two monosaccharides, is called a disaccharide and is conjoined together by a glycosidic or ether bond. The reverse reaction can also occur, using a molecule of water to split up a disaccharide and break the glycosidic bond; this is termed hydrolysis. The most well-known disaccharide is sucrose, ordinary sugar (in scientific contexts, called table sugar or cane sugar to differentiate it from other sugars). Sucrose consists of a glucose molecule and a fructose molecule joined together. Another important disaccharide is lactose, consisting of a glucose molecule and a galactose molecule. As most humans age, the production of lactase, the enzyme that hydrolyzes lactose back into glucose and galactose, typically decreases. This results in lactase deficiency, also called lactose intolerance. Sugar polymers are characterised by having reducing or non-reducing ends. A reducing end of a carbohydrate is a carbon atom which can be in equilibrium with the open-chain aldehyde or keto form. If the joining of monomers takes place at such a carbon atom, the free hydroxy group of the pyranose or furanose form is exchanged with an OH-side chain of another sugar, yielding a full acetal. This prevents opening of the chain to the aldehyde or keto form and renders the modified residue non-reducing. Lactose contains a reducing end at its glucose moiety, whereas the galactose moiety form a full acetal with the C4-OH group of glucose. Saccharose does not have a reducing end because of full acetal formation between the aldehyde carbon of glucose (C1) and the keto carbon of fructose (C2). When a few (around three to six) monosaccharides are joined together, it is called an oligosaccharide (oligo- meaning "few"). These molecules tend to be used as markers and signals, as well as having some other uses. Many monosaccharides joined together make a polysaccharide. They can be joined together in one long linear chain, or they may be branched. Two of the most common polysaccharides are cellulose and glycogen, both consisting of repeating glucose monomers. Cellulose is made by plants and is an important structural component of their cell walls. Humans can neither manufacture nor digest it. Glycogen, on the other hand, is an animal carbohydrate; humans use it as a form of energy storage. Glucose is the major energy source in most life forms; a number of catabolic pathways converge on glucose. For instance, polysaccharides are broken down into their monomers (glycogen phosphorylase removes glucose residues from glycogen). Disaccharides like lactose or sucrose are cleaved into their two component monosaccharides. Glucose is mainly metabolized by a very important and ancient ten-step pathway called glycolysis, the net result of which is to break down one molecule of glucose into two molecules of pyruvate; this also produces a net two molecules of ATP, the energy currency of cells, along with two reducing equivalents in the form of converting NAD+ to NADH. This does not require oxygen; if no oxygen is available (or the cell cannot use oxygen), the NAD is restored by converting the pyruvate to lactate (e. g. in humans) or to ethanol plus carbon dioxide (e. g. in yeast). Other monosaccharides like galactose and fructose can be converted into intermediates of the glycolytic pathway. In aerobic cells with sufficient oxygen, like most human cells, the pyruvate is further metabolized. It is irreversibly converted to acetyl-CoA, giving off one carbon atom as the waste product carbon dioxide, generating another reducing equivalent as NADH. The two molecules acetyl-CoA (from one molecule of glucose) then enter the citric acid cycle, producing two more molecules of ATP, six more NADH molecules and two reduced (ubi)quinones (via FADH2 as enzyme-bound cofactor), and releasing the remaining carbon atoms as carbon dioxide. The produced NADH and quinol molecules then feed into the enzyme complexes of the respiratory chain, an electron transport system transferring the electrons ultimately to oxygen and conserving the released energy in the form of a proton gradient over a membrane (inner mitochondrial membrane in eukaryotes). Thereby, oxygen is reduced to water and the original electron acceptors NAD+ and quinone are regenerated. This is why humans breathe in oxygen and breathe out carbon dioxide. The energy released from transferring the electrons from high-energy states in NADH and quinol is conserved first as proton gradient and converted to ATP via ATP synthase. This generates an additional 28 molecules of ATP (24 from the 8 NADH + 4 from the 2 quinols), totaling to 32 molecules of ATP conserved per degraded glucose (two from glycolysis + two from the citrate cycle). It is clear that using oxygen to completely oxidize glucose provides an organism with far more energy than any oxygen-independent metabolic feature, and this is thought to be the reason why complex life appeared only after Earth's atmosphere accumulated large amounts of oxygen. In vertebrates, vigorously contracting skeletal muscles (during weightlifting or sprinting, for example) do not receive enough oxygen to meet the energy demand, and so they shift to anaerobic metabolism, converting glucose to lactate (lactic acid). The liver regenerates the glucose, using a process called gluconeogenesis. This process is not quite the opposite of glycolysis, and actually requires three times the amount of energy gained from glycolysis (six molecules of ATP are used, compared to the two gained in glycolysis). Analogous to the above reactions, the glucose produced can then undergo glycolysis in tissues that need energy, be stored as glycogen (or starch in plants), or be converted to other monosaccharides or joined into di- or oligosaccharides. Carbohydrates are molecules that contain oxygen, hydrogen, and carbon atoms. They may also contain other elements such as sulfur or nitrogen, but these are usually minor components. They consist of monosaccharide sugars, of varying chain lengths, that have the general chemical formula Cn(H2O)n or are derivatives of such.The smallest value for n is 3. A 3-carbon sugar is referred to as a triose, whereas a 6-carbon sugar is called a hexose (see monosaccharides below). Certain carbohydrates are important for storing and transporting energy in most organisms, including plants and animals, and are major structural elements in many organisms (eg cellulose in plants). In addition they play major roles in cell to cell communication, the immune system, fertilization, pathogenesis, blood clotting, and development. Carbohydrates can be classified by the number of constituent sugar units: monosaccharides (such as glucose and fructose), disaccharides (such as sucrose and lactose), oligosaccharides, and polysaccharides (such as starch, glycogen, and cellulose). Glucose as a straight-chain carbohydrate (Fischer projection) Fructose (Fischer projection) Pure carbohydrates contain carbon, hydrogen, and oxygen atoms, in a 1:2:1 molar ratio, giving the general formula Cn(H2O)n. (This applies only to monosaccharides, see below, although all carbohydrates have the more general formula Cn(H2O)m.) However, many important carbohydrates deviate from this, such as deoxyribose and glycerol. Sometimes compounds containing other elements are also counted as carbohydrates (e.g. glucosamine and chitin, which contain nitrogen). The simplest carbohydrates are monosaccharides, which are small straight-chain aldehydes and ketones with many hydroxyl groups added, usually one on each carbon except the functional group. Other carbohydrates are composed of monosaccharide units and break down under hydrolysis. These may be classified as disaccharides, oligosaccharides, or polysaccharides, depending on whether they have two, several, or many monosaccharide units. Contents [hide] 1 Monosaccharides 2 Disaccharides 3 Oligosaccharides and polysaccharides 4 Nutrition 4.1 Foods that are high in carbohydrates 4.2 Classification 5 Catabolism 6 Anabolism 7 See also 8 References 9 External links Monosaccharides Monosaccharides may be divided into aldoses, which have an aldehyde group on the first carbon atom, and ketoses, which typically have a ketone group on the second. They may also be divided into trioses, tetroses, pentoses, hexoses,and many more. This all depends on how many carbon atoms they contain. For instance, glucose is an aldohexose, fructose a ketohexose, and ribose an aldopentose. Further, each carbon atom that supports a hydroxyl group (except for the first and last) is a stereogenic centre, allowing a number of different enantiomers and stereoisomers for carbohydrates with the same basic structure. For instance, galactose is an aldohexose but has different properties from glucose because the atoms are arranged differently. A heterocyclic form of ribose (Haworth projection) The straight-chain structure described here is only one of the forms a monosaccharide may take. The aldehyde or ketone group may react with a hydroxyl group on a different carbon atom to form a hemiacetal or hemiketal, in which case there is an oxygen bridge between the two carbon atoms, forming a heterocyclic ring. Rings with five and six atoms are called furanose and pyranose forms and exist in equilibrium with the straight-chain form. It should be noted that the ring form has one more stereogenic centre than the straight-chain form, and so has both an alpha and a beta form, which interconvert in equilibrium. However, the carbohydrate may further react with an alcohol to form an acetal or ketal, in which case the two forms become distinct. This is the basic type of link between the monosaccharide units of larger carbohydrates. Excessive consumption may cause obesity. An aldose is a monosaccharide (a certain type of sugar) containing one aldehyde group per molecule and having a chemical formula of the form CnH2nOn (n>=3). Triose: glyceraldehyde Tetroses: erythrose, threose Pentoses: ribose, arabinose, xylose, lyxose Hexoses: allose, altrose, glucose, mannose, gulose, idose, galactose, talose With only 3 carbon atoms, glyceraldehyde is the simplest of all aldoses. Aldoses isomerize to ketoses in the Lobry-de Bruyn-van Ekenstein transformation An aldehyde is an organic compound containing a terminal carbonyl group. This functional group, which consists of a carbon atom which is bonded to a hydrogen atom and double-bonded to an oxygen atom (chemical formula -CHO), is called the aldehyde group. The aldehyde group is also called the formyl or methanoyl group. The word aldehyde seems to have arisen from alcohol dehydrogenated. In the past, aldehydes were sometimes named after the corresponding alcohols, for example vinous aldehyde for acetaldehyde. (Vinous is from Latin vinum = wine, the traditional source of ethanol; compare vinyl.) The aldehyde group is polar. Oxygen, being more electronegative, pulls the electrons in the carbon-oxygen bond towards itself, thus creating an electron deficiency at the carbon atom. Disaccharides Main article: Disaccharide Disaccharides are composed of two monosaccharide units bound together by a covalent glycosidic bond. The binding between the two sugars results in the loss of a hydrogen atom (H) from one molecule and a hydroxyl group (OH) from the other. The most common disaccharides are sucrose (cane or beet sugar - made from one glucose and one fructose), lactose (milk sugar - made from one glucose and one galactose), maltose (made of two glucoses linked alpha-1,4) and cellobiose (made of two glucoses linked beta-1,4). The formula of these disaccharides is C12H22O11. Other examples of disaccharides include trehalose, chitobiose, laminaribiose, kojibiose and xylobiose. [edit] Oligosaccharides and polysaccharides Main articles: Oligosaccharide and Polysaccharide Oligosaccharides and polysaccharides are composed of longer chains of monosaccharide units bound together by glycosidic bonds. The distinction between the two is based upon the number of monosaccharide units present in the chain. Oligosaccharides typically contain between two and nine monosaccharide units, and polysaccharides contain greater than ten monosaccharide units. Definitions of how large a carbohydrate must be to fall into each category vary according to personal opinion. Examples of oligosaccharides include the disaccharides mentioned above, the trisaccharide raffinose and the tetrasaccharide stachyose. Oligosaccharides are found as a common form of protein posttranslational modification. Such posttranslational modifications include the Lewis oligosaccharides responsible for blood group incompatibilities, the alpha-Gal epitope responsible for hyperacute rejection in xenotransplanation, and O-GlcNAc modifications. Polysaccharides represent an important class of biological polymer. Examples include starch, cellulose, chitin, glycogen, callose, laminarin, xylan, and galactomannan. [edit] Nutrition Unrefined grain products are rich sources of complex carbohydrates Carbohydrates require less water to digest than proteins or fats and are the most common source of energy. Proteins and fat are vital building components for body tissue and cells, and thus it could be considered advisable not to deplete such resources by necessitating their use in energy production. Carbohydrates, like proteins, contain 4 kilocalories per gram while fats contain 9 kilocalories and alcohol contains 7 kilocalories per gram. Based on evidence for risk of heart disease and obesity, the Institute of Medicine recommends that American and Canadian adults get between 40-65% of dietary energy from carbohydrates.
[1] The Food and Agriculture Organization and World Health Organization jointly recommend that national dietary guidelines set a goal of 55-75% of total energy from carbohydrates.
[2] Foods that are high in carbohydrates Breads, pastas, beans, potatoes, bran, rice and cereals are all high in carbohydrates

All about amioacids

New a a Amino acids play central roles both as building blocks of proteins and as intermediates in metabolism. The 20 amino acids that are found within proteins convey a vast array of chemical versatility. The precise amino acid content, and the sequence of those amino acids, of a specific protein, is determined by the sequence of the bases in the gene that encodes that protein. The chemical properties of the amino acids of proteins determine the biological activity of the protein. Proteins not only catalyze all (or most) of the reactions in living cells, they control virtually all cellular process. In addition, proteins contain within their amino acid sequences the necessary information to determine how that protein will fold into a three dimensional structure, and the stability of the resulting structure. The field of protein folding and stability has been a critically important area of research for years, and remains today one of the great unsolved mysteries. It is, however, being actively investigated, and progress is being made every day. As we learn about amino acids, it is important to keep in mind that one of the more important reasons to understand amino acid structure and properties is to be able to understand protein structure and properties. We will see that the vastly complex characteristics of even a small, relatively simple, protein are a composite of the properties of the amino acids which comprise the protein. Humans can produce 10 of the 20 amino acids. The others must be supplied in the food. Failure to obtain enough of even 1 of the 10 essential amino acids, those that we cannot make, results in degradation of the body's proteins—muscle and so forth—to obtain the one amino acid that is needed. Unlike fat and starch, the human body does not store excess amino acids for later use—the amino acids must be in the food every day. The 10 amino acids that we can produce are alanine, asparagine, aspartic acid, cysteine, glutamic acid, glutamine, glycine, proline, serine and tyrosine. Tyrosine is produced from phenylalanine, so if the diet is deficient in phenylalanine, tyrosine will be required as well. The essential amino acids are arginine (required for the young, but not for adults), histidine, isoleucine, leucine, lysine, methionine, phenylalanine, threonine, tryptophan, and valine. These amino acids are required in the diet. Plants, of course, must be able to make all the amino acids. Humans, on the other hand, do not have all the the enzymes required for the biosynthesis of all of the amino acids. Why learn these structures and properties? Aliphatic Amino Acids Aliphatic R groups are nonpolar and hydrophobic. Hydrophobicity increases with increasing number of C atoms in the hydrocarbon chain. Although these amino acids prefer to remain inside protein molecules, alanine and glycine are ambivalent, meaning that they can be inside or outside the protein molecule. Glycine has such a small side chain that it does not have much effect on the hydrophobic interactions. The structures below are shown in the ionization state that predominates at pH 7. Less hydrophobic More hydrophobic Aromatic Amino Acids Aromatic amino acids are relatively nonpolar. To different degrees, all aromatic amino acids absorb ultraviolet light. Tyrosine and tryptophan absorb more than do phenylalanine; tryptophan is responsible for most of the absorbance of ultraviolet light (ca. 280 nm) by proteins. Tyrosine is the only one of the aromatic amino acids with an ionizable side chain. Tyrosine is one of three hydroxyl containing amino acids. Least hydrophobic Very hydrophobic Acidic Amino Acids and their Amides Acidic amino acids are polar and negatively charged at physiological pH. Both acidic amino acids have a second carboxyl group. Amides are polar and uncharged, and not ionizable. All are very hydrophilic. Acidic amino acid Amide Acidic amino acid Amide Basic Amino Acids Basic amino acids are polar and positively charged at pH values below their pKa's, and are very hydrophilic. Even though the basic amino acids are almost always in contact with the solvent, the side chain of lysine has a marked hydrocarbon character, so it is often found NEAR the surface, with the amino group of the side chain in contact with solvent. Note that in the drawing, histidine is shown in the protonated form, while at pH 7.0, the imidazole would exist predominantly in the neutral form. Cyclic Amino Acid Proline is the only cyclic amino acid. It is nonpolar and shares many properties with the aliphatic group. Proline is one of the ambivalent amino acids, meaning that it can be inside or outside of a protein molecule. Due to its unique structure, proline occurs in proteins frequently in turns or bends, which are often on the surface. The structure shown is of the amino acid in the ionization state that predominates at pH 7.0. Hydroxyl Amino Acids Hydroxyl amino acids are polar, uncharged at physiological pH, and hydrophilic. The phenolic hydroxyl ionizes with a pKa of 10 to yield the phenolate anion. The hydroxyl groups of serine and threonine are so high that they are generally regarded as nonionizing. Sulfur-Containing Amino Acids The sulfur-containing amino acids (cysteine and methionine) are generally considered to be nonpolar and hydrophobic. In fact, methionine is one of the most hydrophobic amino acids and is almost always found on the interior of proteins. Cysteine on the other hand does ionize to yield the thiolate anion. Even so, it is uncommon to find cysteine on the surface of a protein. There are several reasons. First, sulfur has a low propensity to hydrogen bond, unlike oxygen. A consequence of this fact is that H2S is a gas under conditions that H2O is a liquid. Second, the thiol group of cysteine can react with other thiol groups in an oxidation reaction that yields a disulfide bond. Perhaps as a consequence, cysteine residues are most frequently buried inside proteins. Individua amino acids Alanine A (Ala) Chemical Properties: Aliphatic(Aliphatic R-group) Physical Properties: Nonpolar Alanine is a hydrophobic molecule. It is ambivalent, meaning that it can be inside or outside of the protein molecule. The α carbon of alanine is optically active; in proteins, only the L-isomer is found. Note that alanine is the α-amino acid analog of the α-keto acid pyruvate, an intermediate in sugar metabolism. Alanine and pyruvate are interchangeable by a transamination reaction. Interchangeable with Pyruvate Chemical Properties: Basic (Basic R-group) Physical Properties: Polar (positively charged) Arginine, an essential amino acid, has a positively charged guanidino group. Arginine is well designed to bind the phosphate anion, and is often found in the active centers of proteins that bind phosphorylated substrates. As a cation, arginine, as well as lysine, plays a role in maintaining the overall charge balance of a protein. Arginine also plays an important role in nitrogen metabolism. In the urea cycle, the enzyme arginase cleaves (hydrolyzes) the guanidinium group to yield urea and the L-amino acid ornithine. Ornithine is lysine with one fewer methylene groups in the side chain. L-ornithine is not normally found in proteins. There are 6 codons in the genetic code for arginine, yet, although this large a number of codons is normally associated with a high frequency of the particular amino acid in proteins, arginine is one of the least frequent amino acids. The discrepancy between the frequency of the amino acid in proteins and the number of codons is greater for arginine than for any other amino acid. Asparagine N (Asn) Chemical Properties: Neutral (Amides of acidic amino acids R-group) Physical Properties: Polar (uncharged) Asparagine is the amide of aspartic acid. The amide group does not carry a formal charge under any biologically relevant pH conditions. The amide is rather easily hydrolyzed, converting asparagine to aspartic acid. This process is thought to be one of the factors related to the molecular basis of aging. Asparagine has a high propensity to hydrogen bond, since the amide group can accept two and donate two hydrogen bonds. It is found on the surface as well as buried within proteins. Asparagine is a common site for attachment of carbohydrates in glycoproteins. Cysteine C (Cys) Chemical Properties: Sulfur-containing (Sulfur containing group) Physical Properties: Polar (uncharged) Cysteine is one of two sulfur-containing amino acids; the other is methionine. Cysteine differs from serine in a single atom-- the sulfur of the thiol replaces the oxygen of the alcohol. The amino acids are, however, much more different in their physical and chemical properties than their similarity might suggest. Consider, for example, the differences between H2O and H2S. The hydrogen bonding propensity of water is well known and is responsible for many of its remarkable features. Under similar conditions of temperature and pressure, however, H2S is a gas as a consequence of its weak H-bonding propensity. Furthermore, the proton of the thiol of cysteine is much more acid than the hydroxylic proton of serine, making the nucleophilic thiol(ate) much more reactive than the hydroxyl of serine. Cysteine also plays a key role in stabilizing extracellular proteins. Cysteine can react with itself to form an oxidized dimer by formation of a disulfide bond. The environment within a cell is too strongly reducing for disulfides to form, but in the extracellular environment, disulfides can form and play a key role in stabilizing many such proteins, such as the digestive enzymes of the small intestine. Cysteine and methionine are the only sulfur-containing amino acids. Glutamic Acid E (Glu) Chemical Properties: Acidic (Acidic R-group and their amides) Physical Properties: Polar (charged) Interconvertible with α-ketoglutarate Biosynthesis of Proline Glutamic acid has one additional methylene group in its side chain than does aspartic acid. The side chain carboxyl of aspartic acid is referred to as the β carboxyl group, while that of glutamic acid is referred to as the γ carboxyl group. The pKa of the γ carboxyl group for glutamic acid in a polypeptide is about 4.3, significantly higher than that of aspartic acid. This is due to the inductive effect of the additional methylene group. In some proteins, due to a vitamin K dependent carboxylase, some glutamic acids will be dicarboxylic acids, referred to as γ carboxyglutamic acid, that form tight binding sites for calcium ion. Glutamic acid is interconvertible by transamination withα-ketoglutarate Glutamic acid and α-ketoglutarate, an intermediate in the Krebs cycle, are interconvertible by transamination. Glutamic acid can therefore enter the Krebs cycle for energy metabolism, and be converted by the enzyme glutamine synthetase into glutamine, which is one of the key players in nitrogen metabolism. Biosynthesis of Proline Note also that glutamic acid is easily converted into proline. First, the γ carboxyl group is reduced to the aldehyde, yielding glutamate semialdehyde. The aldehyde then reacts with the α-amino group, eliminating water as it forms the Schiff base. In a second reduction step, the Schiff base is reduced, yielding proline.

INTRODUCTION TO BIOCHEMISTRY

INTRODUCTION TO BIOCHEMISTRY The study of the chemical substances and vital processes occurring in living organisms. The chemical composition of a particular living system or biological substance. Biochemistry is the application of chemistry to the study of biological processes at the cellular and molecular level. It emerged as a distinct discipline around the beginning of the 20th century when scientists combined chemistry, physiology and biology to investigate the chemistry of living systems. Biochemistry has become the foundation for understanding all biological processes. It has provided explanations for the causes of many diseases in humans, animals and plants. It can frequently suggest ways by which such diseases may be treated or cured. Biochemists study how living organisms extract food and energy from their environment and how they use the extracted molecules to make more of themselves. Buchner, by taking apart yeast cells, had opened the way to ask biochemical questions like: What kinds of molecules cause fermentation? How many different molecules are necessary? Why does the yeast cell do it? Why does it only happen if you keep oxygen out? These are questions that can be answered by separating the "dissolved substances" in the "juice" and asking what they are, how they interact with each other, and how their properties are related to their chemical nature.By using this approach, biochemists have succeeded in...Discovering that although too much cholesterol can cause heart disease, our bodies make cholesterol because it is an essential component of the membranes of our cells.Finding that cells distinctively mark themselves by putting specific groups of sugars, linked together in recognizable patterns, on their surfaces. Your body will reject transplanted tissue if the cells of that tissue have the wrong pattern of sugar groups on their surfaces.Learning that one of the reasons plants require the mineral nutrient magnesium is because it forms part of the structure of chlorophyll, the molecule plants use to trap solar energy.Exploring the way penicillin kills bacteria to discover that it prevents them from putting together the chemical structure of their cell walls. Biochemistry focuses on the study of life at molecular level - how genes and proteins regulate cells, tissues, organs and ultimately whole organisms like you. As you go about your daily life biochemistry is sure to be involved. It has a role in: understanding the causes of diseases; use of engineered therapeutic proteins in medicine; food production; understanding how cells function. This means you can see the effects of biochemistry all around you! It is central to all areas of the Biological or Life Sciences. The aim is to provide an understanding of every aspect of the structure and function of living things at the molecular level. It is a practical laboratory science that applies the molecular approaches of chemistry to the vast variety of biological systems.Biochemists work at all levels and with all types of biological organisms, ranging from biomolecules to man. There are close links with other specialist life sciences, such as Cell Biology, Genetics, Microbiology, Molecular Biology, Physiology and Pharmacology. In fact, in many cases the distinctions between these disciplines are becoming increasingly blurred. They use biochemical techniques and biochemists work in all these areas. Biochemistry offers the tremendous challenge of seeking to understand the most fundamental of life's processes at the molecular level, and to utilise this knowledge for the benefit of mankind. You will have read, for example, how biochemists, working with colleagues in other disciplines, have developed the new technologies of Molecular Biology and Genetic Engineering. These have enabled the production of therapeutically important human proteins such as insulin and blood clotting factors by cloning procedures, thus avoiding costly, time-consuming and inefficient isolation of these molecules from biological sources; the identification and possible remedying of genetic problems; and the use of DNA fingerprinting in forensic science. Biochemistry is the study of chemical processes associated with living organisms. Biochemists use concepts of biology, chemistry, physics, mathematics, microbiology, and genetics to unravel the complex puzzles of life. Biochemical techniques are used in clinical diagnosis of infectious diseases, genetic disorders, and cancer; as well as in many forms of research to improve the quality of our lives. Biochemists identify biological problems then develop and apply appropriate techniques to solve them at the molecular levelBiochemists study the most basic of life processes; for example, identifying the way in which DNA, which carries the genetic information, is transferred between cells and can be manipulated. This has led to the development of new technologies such as Molecular Biology and Genetic Engineering. The resulting recombinant DNA technology has formed the basis of modern biotechnology (e.g. production of human insulin), medical developments (e.g. prenatal diagnosis and genetic counselling) and forensic science (e.g. DNA fingerprinting).DNA directs the production of proteins. These have diverse functions, such as catalysing biological reactions (enzymes), carrying oxygen round the body (haemoglobin), protecting us from infection (antibodies) and holding us together (collagen). Using both simple and high-technology methods, biochemists work out how these proteins function. Biochemists also develop methods for making use of proteins, such as enzymes in biotechnology and antibodies in hormone analysis.With knowledge of the basic molecular mechanisms, biochemists study how life processes are integrated to allow individual cells to function and interact to form complex organisms. They work with all sorts of organisms, from viruses and bacteria to plants and man.These are just a few of the areas. It would take a whole book, in fact many books, to do justice to the multitude of roles of biochemists. Biochemists work in many walks of life - in industry, hospitals, agriculture, research institutes, education and associated areas. There are many areas of everyday life as diverse as medical products and diagnostics, new food and its safety, crop improvement, cosmetics and forensic science that owe their development or even existence to biochemists IndustryPharmaceutical, food, brewing, biotechnology and agrochemical companies all need and employ biochemists to develop new products and to monitor the production, quality control and safety of existing ones. MedicineHospitals, public health laboratories and medical research institutes, as well as the pharmaceutical industry, all require biochemists. Here they provide a diagnostic service, carrying out tests on blood, urine and other body fluids, alongside researching the underlying causes of disease and the methods of treatment. Agriculture and the EnvironmentBiochemists and biotechnologists, who often have a biochemistry degree, working in agriculture have been responsible for many developments, such as pest-resistant crops, improvements in crop yields and tomatoes that keep better. They also monitor the environment. Employers include seed companies, local government, the Civil Service and water authorities. EducationAll levels of education offer prospects for biochemists. The combination of biology and chemistry, along with the training in numerical and analytical skills that is given in any area of science, makes biochemistry ideal for teaching throughout the school age range. There are also opportunities for more advanced teaching, usually associated with research, in universities and colleges, and medical, dental and veterinary schools. Away from ScienceA science background can be an excellent starting point for many other careers. Biochemistry is a numerate subject that develops analytical thinking, creativity in problem solving, and the ability to handle large amounts of complex information - skills required in jobs in all walks of life including, for example, sales and marketing, accountancy and finance, journalism, and patent work. Biochemists have become successful popular authors and even a national president! Originally, it was generally believed that life was not subject to the laws of science the way non-life was. It was thought that only living beings could produce the molecules of life (from other, previously existing biomolecules). Then, in 1828, Friedrich Wöhler published a paper about the synthesis of urea, proving that organic compounds can be created artificially. The dawn of biochemistry may have been the discovery of the first enzyme, diastase (today called amylase), in 1833 by Anselme Payen. Eduard Buchner contributed the first demonstration of a complex biochemical process outside of a cell in 1896: alcoholic fermentation in cell extracts of yeast. Although the term “biochemistry” seems to have been first used in 1881, it is generally accepted that the formal coinage of biochemistry occurred in 1903 by Carl Neuber, a German chemist. Since then, biochemistry has advanced, especially since the mid-20th century, with the development of new techniques such as chromatography, X-ray diffraction, NMR spectroscopy, radioisotopic labelling, electron microscopy and molecular dynamics simulations. These techniques allowed for the discovery and detailed analysis of many molecules and metabolic pathways of the cell, such as glycolysis and the Krebs cycle (citric acid cycle). Today, the findings of biochemistry are used in many areas, from genetics to molecular biology and from agriculture to medicine macromolecule is a large molecule with a large molecular mass, but generally the use of the term is restricted to polymers and molecules which structurally include polymers. [1] Illustration of a polypeptide macromolecule Many examples come from biology and in particular biochemistry. In case of "biomacromolecules" or biopolymers, there are proteins, carbohydrates and nucleic acids (such as DNA). Lipids (fat) are not considered true macromolecules by most biologists as they are not covalently bonded, and so are not true polymers. Synthetic examples include plastics. The integral domains of crystals and metals, while composed of very large numbers of atoms joined by molecule-like bonds, are rarely referred to as "macromolecules." The term macromolecule is also sometimes used to refer to aggregates of two or more macromolecules held together by intermolecular forces rather than by chemical "bonds". This usage is common in particular when the individual macromolecules involved aggregate or "assemble" spontaneously and rarely exist in isolation. Such an aggregate is more properly called a macromolecular complex. In such a context, individual macromolecules are often referred to as subunits (see e.g. protein subunit). Substances that are composed of macromolecules often have unusual physical properties. The properties of liquid crystals and such elastomers as rubber are examples. Although too small to see, individual pieces of DNA in solution can be broken in two simply by suctioning the solution through an ordinary straw. This is not true of smaller molecules. The 1964 edition of Linus Pauling's College Chemistry asserted that DNA in nature is never longer than about 5000 base pairs. This is because biochemists were inadvertently and with perfect consistency breaking their samples into pieces. In fact, the DNA of chromosomes can be tens of millions of base pairs long. Another common macromolecular property that does not characterize smaller molecules is the need for assistance in dissolving into solution. Many require salts or particular ions to dissolve in water. Proteins will denature if the solute concentration of their solution is too high or too low. According to IUPAC recommendations the term macromolecule is reserved for an individual molecule, and the term polymer is used as to denote a substance composed of macromolecules. Polymer may also be employed unambiguously as an adjective, according to accepted usage, e.g. polymer blend, polymer molecule. [2] [edit] External links Major categories of bio-compounds: Carbohydrates : sugar -- disaccharide -- polysaccharide -- cholesterol -- starch -- glycogen Lipids : fatty acid -- fats -- essential oils -- oils -- waxes Nucleic acids : DNA -- RNA -- mRNA -- tRNA -- rRNA -- codon -- adenosine -- cytosine -- guanine -- thymine -- uracil Proteins : amino acid -- glycine -- arginine -- lysine peptide -- primary structure -- secondary structure -- tertiary structure -- conformation -- protein folding Chemical properties: molecular bond -- covalent bond -- ionic bond -- hydrogen bond -- ester -- ethyl molecular charge -- hydrophilic -- hydrophobic -- polar pH -- acid -- alkaline -- base oxidation -- reduction -- hydrolysis Structural compounds: In cells: flagellin -- peptidoglycan -- myelin -- actin -- myosin In animals: chitin -- keratin -- collagen -- silk In plants: cellulose -- lignin -- cell wall Enzymes and enzyme activity: enzyme kinetics -- enzyme inhibition proteolysis -- ubiquitin -- proteasome kinase -- dehydrogenase Membranes : fluid mosaic model -- diffusion -- osmosis phospholipids -- glycolipid -- glycocalyx -- antigen -- isoprene ion channel -- proton pump -- electron transport -- ion gradient -- antiporter -- symporter -- quinone -- riboflavin Energy pathways : pigments : chlorophyll -- carotenoids -- xanthophyll -- cytochrome -- phycobilin -- bacteriorhodopsin -- hemoglobin -- myoglobin -- absorption spectrum -- action spectrum -- fluorescence Photosynthesis : light reaction -- dark reaction Fermentation : Acetyl-CoA -- lactic acid Cellular respiration : Adenosine triphosphate (ATP) -- NADH -- pyruvate -- oxalate -- citrate Chemosynthesis Regulation hormones : auxin signal transduction -- growth factor -- transcription factor -- protein kinase -- SH3 domain Malfunctions : tumor -- oncogene -- tumor suppressor gene Receptors : Integrin -- transmembrane receptor -- ion channel Techniques : electrophoresis -- chromatography -- mass spectrometry -- x-ray diffraction -- Southern blot -- fractionation -- Gram stain Objective To describe functional groups of organic compounds of biological interest and provide some examples of chemical reactions and interconversions among these. Reference: Stryer, 4th edition, 1994, Chapter 1, pp. 3-16 I. Some Examples of Chemical Reactions I. Functional Groups a. Alcohols R-CH2-OH(primary); R2-CH-OH (secondary); R3-C-OH (tertiary) Sustitution on the Carbon atom defines whether the alcohol is primary, secondary or tertiary. Carbon atom in these cases is in sp3 hybrid state. b. Aldehydes and Ketones R-CHO; R2-C=O Carbon atom is double bonded to oxygen in both cases. The difference is in substitutions. Aldehyde has an R group and a H atom whereas in a ketone both substituents are R groups. Carbon atom in these cases is in sp2 hybrid state. c. Acids R-C=O Acids are able to dissociate into H+ and anions ! OH d. Acid Anhydrides (R-CO)2O Acid anhydrides are formed from molecules of same or different acids with the elimination of a molecule of water. An example would the formation of pyrophosphate from two molecules of phosphoric acid. e. Esters RCH2-O-COCH3 An ester is formed from an alcohol and an acid with the elimination of a molecule of water.Physiological examples include formation of triacyl glycerols from fatty acids and glycerol. f. Unsaturated Compounds Unsaturated compouds like R-CH=CH---- are formed either by the elimination of water fron a hydroxy compund like R-CHOH-CH2---- or by dehydrogenation of compunds like R-CH2-CH2---. The double bond may be cis or trans depending on the positions of hydrogen atoms in space. In cis configuration, H atoms would lie in a plane perpendicular to the plane of the double bond whereas iin trans configuration H atoms lie in the same plane as the double bond. g. Amines and Amides II. Reaction Types A. Oxidation/Reduction (removal of electron or reaction with O2) B. Esterification (carboxylic acid plus alcohol) C. Hydrolysis (cleavage of a bond by water) D. Phosphorolysis (cleavage of a bond by inorganic phosphate) E. Decarboxylation F. Deamination G. Transamination (amino group transfer) H. Phosphorylation (ester bond on sugars, some amino acids, bases) I. Dehydration J. Phosphorylation (transfer of a phosphate group) K. Transmethylation L. Condensation III. Coupled Reactions Decarboxylation and oxidation/reduction Study Assignment I Draw the structures of molecules that have the functional groups listed above and identify the functional groups. Some molecules will have several functional groups. Study Assignment II For reaction types listed in IIA-L, above, write out the entire reaction, including molecular structure. Use material presented to you in Biochemistry lectures to complete the assignment.

Basics of amino acids

New a a
Amino acids play central roles both as building blocks of proteins and as intermediates in metabolism. The 20 amino acids that are found within proteins convey a vast array of chemical versatility. The precise amino acid content, and the sequence of those amino acids, of a specific protein, is determined by the sequence of the bases in the gene that encodes that protein. The chemical properties of the amino acids of proteins determine the biological activity of the protein. Proteins not only catalyze all (or most) of the reactions in living cells, they control virtually all cellular process. In addition, proteins contain within their amino acid sequences the necessary information to determine how that protein will fold into a three dimensional structure, and the stability of the resulting structure. The field of protein folding and stability has been a critically important area of research for years, and remains today one of the great unsolved mysteries. It is, however, being actively investigated, and progress is being made every day.
As we learn about amino acids, it is important to keep in mind that one of the more important reasons to understand amino acid structure and properties is to be able to understand protein structure and properties. We will see that the vastly complex characteristics of even a small, relatively simple, protein are a composite of the properties of the amino acids which comprise the protein.
Humans can produce 10 of the 20 amino acids. The others must be supplied in the food. Failure to obtain enough of even 1 of the 10 essential amino acids, those that we cannot make, results in degradation of the body's proteins—muscle and so forth—to obtain the one amino acid that is needed. Unlike fat and starch, the human body does not store excess amino acids for later use—the amino acids must be in the food every day.
The 10 amino acids that we can produce are alanine, asparagine, aspartic acid, cysteine, glutamic acid, glutamine, glycine, proline, serine and tyrosine. Tyrosine is produced from phenylalanine, so if the diet is deficient in phenylalanine, tyrosine will be required as well. The essential amino acids are arginine (required for the young, but not for adults), histidine, isoleucine, leucine, lysine, methionine, phenylalanine, threonine, tryptophan, and valine. These amino acids are required in the diet. Plants, of course, must be able to make all the amino acids. Humans, on the other hand, do not have all the the enzymes required for the biosynthesis of all of the amino acids.
Why learn these structures and properties?


Aliphatic Amino Acids
Aliphatic R groups are nonpolar and hydrophobic. Hydrophobicity increases with increasing number of C atoms in the hydrocarbon chain. Although these amino acids prefer to remain inside protein molecules, alanine and glycine are ambivalent, meaning that they can be inside or outside the protein molecule. Glycine has such a small side chain that it does not have much effect on the hydrophobic interactions.
The structures below are shown in the ionization state that predominates at pH 7.
Less hydrophobic More hydrophobic
Aromatic Amino Acids
Aromatic amino acids are relatively nonpolar. To different degrees, all aromatic amino acids absorb ultraviolet light. Tyrosine and tryptophan absorb more than do phenylalanine; tryptophan is responsible for most of the absorbance of ultraviolet light (ca. 280 nm) by proteins. Tyrosine is the only one of the aromatic amino acids with an ionizable side chain. Tyrosine is one of three hydroxyl containing amino acids.
Least hydrophobic Very hydrophobic
Acidic Amino Acids and their Amides
Acidic amino acids are polar and negatively charged at physiological pH. Both acidic amino acids have a second carboxyl group.
Amides are polar and uncharged, and not ionizable. All are very hydrophilic.
Acidic amino acid
Amide

Acidic amino acid
Amide


Basic Amino Acids
Basic amino acids are polar and positively charged at pH values below their pKa's, and are very hydrophilic. Even though the basic amino acids are almost always in contact with the solvent, the side chain of lysine has a marked hydrocarbon character, so it is often found NEAR the surface, with the amino group of the side chain in contact with solvent. Note that in the drawing, histidine is shown in the protonated form, while at pH 7.0, the imidazole would exist predominantly in the neutral form.
Cyclic Amino Acid
Proline is the only cyclic amino acid. It is nonpolar and shares many properties with the aliphatic group.
Proline is one of the ambivalent amino acids, meaning that it can be inside or outside of a protein molecule. Due to its unique structure, proline occurs in proteins frequently in turns or bends, which are often on the surface. The structure shown is of the amino acid in the ionization state that predominates at pH 7.0.
Hydroxyl Amino Acids
Hydroxyl amino acids are polar, uncharged at physiological pH, and hydrophilic. The phenolic hydroxyl ionizes with a pKa of 10 to yield the phenolate anion. The hydroxyl groups of serine and threonine are so high that they are generally regarded as nonionizing.

Sulfur-Containing Amino Acids
The sulfur-containing amino acids (cysteine and methionine) are generally considered to be nonpolar and hydrophobic. In fact, methionine is one of the most hydrophobic amino acids and is almost always found on the interior of proteins. Cysteine on the other hand does ionize to yield the thiolate anion. Even so, it is uncommon to find cysteine on the surface of a protein. There are several reasons. First, sulfur has a low propensity to hydrogen bond, unlike oxygen. A consequence of this fact is that H2S is a gas under conditions that H2O is a liquid. Second, the thiol group of cysteine can react with other thiol groups in an oxidation reaction that yields a disulfide bond. Perhaps as a consequence, cysteine residues are most frequently buried inside proteins.


Individua amino acids
Alanine A (Ala)
Chemical Properties: Aliphatic(Aliphatic R-group)
Physical Properties: Nonpolar

Alanine is a hydrophobic molecule. It is ambivalent, meaning that it can be inside or outside of the protein molecule. The α carbon of alanine is optically active; in proteins, only the L-isomer is found.
Note that alanine is the α-amino acid analog of the α-keto acid pyruvate, an intermediate in sugar metabolism. Alanine and pyruvate are interchangeable by a transamination reaction.
Interchangeable with Pyruvate
Chemical Properties: Basic (Basic R-group)
Physical Properties: Polar (positively charged)

Arginine, an essential amino acid, has a positively charged guanidino group. Arginine is well designed to bind the phosphate anion, and is often found in the active centers of proteins that bind phosphorylated substrates. As a cation, arginine, as well as lysine, plays a role in maintaining the overall charge balance of a protein.
Arginine also plays an important role in nitrogen metabolism. In the urea cycle, the enzyme arginase cleaves (hydrolyzes) the guanidinium group to yield urea and the L-amino acid ornithine. Ornithine is lysine with one fewer methylene groups in the side chain. L-ornithine is not normally found in proteins.
There are 6 codons in the genetic code for arginine, yet, although this large a number of codons is normally associated with a high frequency of the particular amino acid in proteins, arginine is one of the least frequent amino acids. The discrepancy between the frequency of the amino acid in proteins and the number of codons is greater for arginine than for any other amino acid.
Asparagine N (Asn)
Chemical Properties: Neutral (Amides of acidic amino acids R-group)
Physical Properties: Polar (uncharged)

Asparagine is the amide of aspartic acid. The amide group does not carry a formal charge under any biologically relevant pH conditions. The amide is rather easily hydrolyzed, converting asparagine to aspartic acid. This process is thought to be one of the factors related to the molecular basis of aging.
Asparagine has a high propensity to hydrogen bond, since the amide group can accept two and donate two hydrogen bonds. It is found on the surface as well as buried within proteins.
Asparagine is a common site for attachment of carbohydrates in glycoproteins.
Cysteine C (Cys)
Chemical Properties: Sulfur-containing
(Sulfur containing group)
Physical Properties: Polar (uncharged)

Cysteine is one of two sulfur-containing amino acids; the other is methionine. Cysteine differs from serine in a single atom-- the sulfur of the thiol replaces the oxygen of the alcohol. The amino acids are, however, much more different in their physical and chemical properties than their similarity might suggest.
Consider, for example, the differences between H2O and H2S. The hydrogen bonding propensity of water is well known and is responsible for many of its remarkable features. Under similar conditions of temperature and pressure, however, H2S is a gas as a consequence of its weak H-bonding propensity. Furthermore, the proton of the thiol of cysteine is much more acid than the hydroxylic proton of serine, making the nucleophilic thiol(ate) much more reactive than the hydroxyl of serine.
Cysteine also plays a key role in stabilizing extracellular proteins. Cysteine can react with itself to form an oxidized dimer by formation of a disulfide bond. The environment within a cell is too strongly reducing for disulfides to form, but in the extracellular environment, disulfides can form and play a key role in stabilizing many such proteins, such as the digestive enzymes of the small intestine.
Cysteine and methionine are the only sulfur-containing amino acids.
Glutamic Acid E (Glu)
Chemical Properties:
Acidic
(Acidic R-group and their amides)
Physical Properties:
Polar (charged)
Interconvertible with α-ketoglutarate Biosynthesis of Proline
Glutamic acid has one additional methylene group in its side chain than does aspartic acid. The side chain carboxyl of aspartic acid is referred to as the β carboxyl group, while that of glutamic acid is referred to as the γ carboxyl group.
The pKa of the γ carboxyl group for glutamic acid in a polypeptide is about 4.3, significantly higher than that of aspartic acid. This is due to the inductive effect of the additional methylene group. In some proteins, due to a vitamin K dependent carboxylase, some glutamic acids will be dicarboxylic acids, referred to as γ carboxyglutamic acid, that form tight binding sites for calcium ion.

Glutamic acid is interconvertible by transamination withα-ketoglutarate
Glutamic acid and α-ketoglutarate, an intermediate in the Krebs cycle, are interconvertible by transamination. Glutamic acid can therefore enter the Krebs cycle for energy metabolism, and be converted by the enzyme glutamine synthetase into glutamine, which is one of the key players in nitrogen metabolism.
Biosynthesis of Proline
Note also that glutamic acid is easily converted into proline. First, the γ carboxyl group is reduced to the aldehyde, yielding glutamate semialdehyde. The aldehyde then reacts with the α-amino group, eliminating water as it forms the Schiff base. In a second reduction step, the Schiff base is reduced, yielding proline.

Major Roles of Biological of Lipids

Biological molecules that are insoluble in aqueous solutions and soluble in organic solvents are classified as lipids. The lipids of physiological importance for humans have four major functions:


1. They serve as structural components of biological membranes.

2. They provide energy reserves, predominantly in the form of triacylglycerols.

3. Both lipids and lipid derivatives serve as vitamins and hormones.

4. Lipophilic bile acids aid in lipid solubilization.
Fatty Acids
Fatty acids fill two major roles in the body:
1. as the components of more complex membrane lipids.
2. as the major components of stored fat in the form of triacylglycerols.
Fatty acids are long-chain hydrocarbon molecules containing a carboxylic acid moiety at one end. The numbering of carbons in fatty acids begins with the carbon of the carboxylate group. At physiological pH, the carboxyl group is readily ionized, rendering a negative charge onto fatty acids in bodily fluids.
Fatty acids that contain no carbon-carbon double bonds are termed saturated fatty acids; those that contain double bonds are unsaturated fatty acids. The numeric designations used for fatty acids come from the number of carbon atoms, followed by the number of sites of unsaturation (eg, palmitic acid is a 16-carbon fatty acid with no unsaturation and is designated by 16:0).

Palmitic Acid
The site of unsaturation in a fatty acid is indicated by the symbol Δ and the number of the first carbon of the double bond (e.g. palmitoleic acid is a 16-carbon fatty acid with one site of unsaturation between carbons 9 and 10, and is designated by 16:1Δ9).
Saturated fatty acids of less than eight carbon atoms are liquid at physiological temperature, whereas those containing more than ten are solid. The presence of double bonds in fatty acids significantly lowers the melting point relative to a saturated fatty acid.
The majority of body fatty acids are acquired in the diet. However, the lipid biosynthetic capacity of the body (fatty acid synthase and other fatty acid modifying enzymes) can supply the body with all the various fatty acid structures needed. Two key exceptions to this are the highly unsaturated fatty acids know as linoleic acid and linolenic acid, containing unsaturation sites beyond carbons 9 and 10. These two fatty acids cannot be synthesized from precursors in the body, and are thus considered the essential fatty acids; essential in the sense that they must be provided in the diet. Since plants are capable of synthesizing linoleic and linolenic acid humans can acquire these fats by consuming a variety of plants or else by eating the meat of animals that have consumed these plant fats.
Physiologically Relevant Fatty Acids
Numerical Symbol Common Name and Struture Comments
14:0 Myristic acid

Often found attached to the N-term. of plasma membrane-associated cytoplasmic proteins
16:0 Palmitic acid

End product of mammalian fatty acid synthesis
16:1Δ9 Palmitoleic acid


18:0 Stearic acid


18:1Δ9 Oleic acid


18:2Δ9,12 Linoleic acid

Essential fatty acid
18:3Δ9,12,15 Linolenic acid

Essential fatty acid
20:4Δ5,8,11,14 Arachidonic acid

Precursor for eicosanoid synthesis


back to the top

Basic Structure of Triacylglycerides
Triacylglycerides are composed of a glycerol backbone to which 3 fatty acids are esterified.

Basic Composition of a Triacylglyceride.

back to the top

Basic Structure of Phospholipids
The basic structure of phospolipids is very similar to that of the triacylglycerides except that C-3 (sn3)of the glycerol backbone is esterified to phosphoric acid. The building block of the phospholipids is phosphatidic acid which results when the X substitution in the basic structure shown in the Figure below is a hydrogen atom. Substitutions include ethanolamine (phosphatidylethanolamine), choline (phosphatidylcholine, also called lecithins), serine (phosphatidylserine), glycerol (phosphatidylglycerol), myo-inositol (phosphatidylinositol, these compounds can have a variety in the numbers of inositol alcohols that are phosphorylated generating polyphosphatidylinositols), and phosphatidylglycerol (diphosphatidylglycerol more commonly known as cardiolipins). See the Lipid Synthesis page for images of the various phospholipids.

Basic Composition of a Phospholipid.
X can be a number of different substituents.

back to the top

Basic Structure of Plasmalogens
Plasmalogens are complex membrane lipids that resemble phospholipids, principally phosphatidylcholine. The major difference is that the fatty acid at C-1 (sn1) of glycerol contains either an O-alkyl (-O-CH2-) or O-alkenyl ether (-O-CH=CH-) species. A basic O-alkenyl ether species is shown in the Figure below where -X can be substituents such as those found in phospholipids described above.

Basic Composition of O-Alkenyl Plasmalogens
One of the most potent alkyl ether plasmalogens is platelet activating factor (PAF: 1-O-1'-enyl-2-acetyl-sn-glycero-3-phosphocholine) which is a choline plasmalogen in which the C-2 (sn2) position of glycerol is esterified with an acetyl group instead of a long chain fatty acid.
PAF functions as a mediator of hypersensitivity, acute inflammatory reactions and anaphylactic shock. PAF is synthesized in response to the formation of antigen-IgE complexes on the surfaces of basophils, neutrophils, eosinophils, macrophages and monocytes. The synthesis and release of PAF from cells leads to platelet aggregation and the release of serotonin from platelets. PAF also produces responses in liver, heart, smooth muscle, and uterine and lung tissues.

structure of PAF


Basic Structure of Sphingolipids
Sphingolipids are composed of a backbone of sphingosine which is derived itself from glycerol. Sphingosine is N-acetylated by a variety of fatty acids generating a family of molecules referred to as ceramides. Sphingolipids predominate in the myelin sheath of nerve fibers. Sphingomyelin is an abundant sphingolipid generated by transfer of the phosphocholine moiety of phosphatidylcholine to a ceramide, thus sphingomyelin is a unique form of a phospholipid.
The other major class of sphingolipids (besides the sphingomyelins) are the glycosphingolipids generated by substitution of carbohydrates to the sn1 carbon of the glycerol backbone of a ceramide. There are 4 major classes of glycosphingolipids:
Cerebrosides: contain a single moiety, principally galactose.
Sulfatides: sulfuric acid esters of galactocerebrosides.
Globosides: contain 2 or more sugars.
Gangliosides: similar to globosides except also contain sialic acid.

Nov 21, 2008

ENZYME KINETICS

Introduction to Enzymes
Enzyme Classifications
Role of Coenzymes
Enzyme Activity Relative to Substrate Type
Enzyme-Substrate Interactions
Chemical Reactions and Rates
Chemical Reaction Order
Enzymes as Biological Catalysts
Michaelis-Menton Kinetics
Inhibition of Enzyme Catalyzed Reactions
Regulation of Enzyme Activity
Allosteric Enzymes
Enzymes in the Diagnosis of Pathology

Enzyme Kinetics by Dr. Peter Birch, University of Paisley
Introduction to Enzymes

Enzymes are biological catalysts responsible for supporting almost all of the chemical reactions that maintain animal homeostasis. Because of their role in maintaining life processes, the assay and pharmacological regulation of enzymes have become key elements in clinical diagnosis and therapeutics. The macromolecular components of almost all enzymes are composed of protein, except for a class of RNA modifying catalysts known as ribozymes. Ribozymes are molecules of ribonucleic acid that catalyze reactions on the phosphodiester bond of other RNAs.

Enzymes are found in all tissues and fluids of the body. Intracellular enzymes catalyze the reactions of metabolic pathways. Plasma membrane enzymes regulate catalysis within cells in response to extracellular signals, and enzymes of the circulatory system are responsible for regulating the clotting of blood. Almost every significant life process is dependent on enzyme activity.
back to the top


--------------------------------------------------------------------------------

Enzyme Classifications

Traditionally, enzymes were simply assigned names by the investigator who discovered the enzyme. As knowledge expanded, systems of enzyme classification became more comprehensive and complex. Currently enzymes are grouped into six functional classes by the International Union of Biochemists (I.U.B.).



Number
Classification
Biochemical Properties

1.
Oxidoreductases
Act on many chemical groupings to add or remove hydrogen atoms.

2.
Transferases
Transfer functional groups between donor and acceptor molecules. Kinases are specialized transferases that regulate metabolism by transferring phosphate from ATP to other molecules.

3.
Hydrolases
Add water across a bond, hydrolyzing it.

4.
Lyases
Add water, ammonia or carbon dioxide across double bonds, or remove these elements to produce double bonds.

5.
Isomerases
Carry out many kinds of isomerization: L to D isomerizations, mutase reactions (shifts of chemical groups) and others.

6.
Ligases
Catalyze reactions in which two chemical groups are joined (or ligated) with the use of energy from ATP.



--------------------------------------------------------------------------------

These rules give each enzyme a unique number. The I.U.B. system also specifies a textual name for each enzyme. The enzyme's name is comprised of the names of the substrate(s), the product(s) and the enzyme's functional class. Because many enzymes, such as alcohol dehydrogenase, are widely known in the scientific community by their common names, the change to I.U.B.-approved nomenclature has been slow. In everyday usage, most enzymes are still called by their common name.

Enzymes are also classified on the basis of their composition. Enzymes composed wholly of protein are known as simple enzymes in contrast to complex enzymes, which are composed of protein plus a relatively small organic molecule. Complex enzymes are also known as holoenzymes. In this terminology the protein component is known as the apoenzyme, while the non-protein component is known as the coenzyme or prosthetic group where prosthetic group describes a complex in which the small organic molecule is bound to the apoenzyme by covalent bonds; when the binding between the apoenzyme and non-protein components is non-covalent, the small organic molecule is called a coenzyme. Many prosthetic groups and coenzymes are water-soluble derivatives of vitamins. It should be noted that the main clinical symptoms of dietary vitamin insufficiency generally arise from the malfunction of enzymes, which lack sufficient cofactors derived from vitamins to maintain homeostasis.

The non-protein component of an enzyme may be as simple as a metal ion or as complex as a small non-protein organic molecule. Enzymes that require a metal in their composition are known as metalloenzymes if they bind and retain their metal atom(s) under all conditions with very high affinity. Those which have a lower affinity for metal ion, but still require the metal ion for activity, are known as metal-activated enzymes.
back to the top


--------------------------------------------------------------------------------

Role of Coenzymes

The functional role of coenzymes is to act as transporters of chemical groups from one reactant to another. The chemical groups carried can be as simple as the hydride ion (H+ + 2e-) carried by NAD or the mole of hydrogen carried by FAD; or they can be even more complex than the amine (-NH2) carried by pyridoxal phosphate.

Since coenzymes are chemically changed as a consequence of enzyme action, it is often useful to consider coenzymes to be a special class of substrates, or second substrates, which are common to many different holoenzymes. In all cases, the coenzymes donate the carried chemical grouping to an acceptor molecule and are thus regenerated to their original form. This regeneration of coenzyme and holoenzyme fulfills the definition of an enzyme as a chemical catalyst, since (unlike the usual substrates, which are used up during the course of a reaction) coenzymes are generally regenerated.
back to the top


--------------------------------------------------------------------------------

Enzyme Relative to Substrate Type

Although enzymes are highly specific for the kind of reaction they catalyze, the same is not always true of substrates they attack. For example, while succinic dehydrogenase (SDH) always catalyzes an oxidation-reduction reaction and its substrate is invariably succinic acid, alcohol dehydrogenase (ADH) always catalyzes oxidation-reduction reactions but attacks a number of different alcohols, ranging from methanol to butanol. Generally, enzymes having broad substrate specificity are most active against one particular substrate. In the case of ADH, ethanol is the preferred substrate.

Enzymes also are generally specific for a particular steric configuration (optical isomer) of a substrate. Enzymes that attack D sugars will not attack the corresponding L isomer. Enzymes that act on L amino acids will not employ the corresponding D optical isomer as a substrate. The enzymes known as racemases provide a striking exception to these generalities; in fact, the role of racemases is to convert D isomers to L isomers and vice versa. Thus racemases attack both D and L forms of their substrate.

As enzymes have a more or less broad range of substrate specificity, it follows that a given substrate may be acted on by a number of different enzymes, each of which uses the same substrate(s) and produces the same product(s). The individual members of a set of enzymes sharing such characteristics are known as isozymes. These are the products of genes that vary only slightly; often, various isozymes of a group are expressed in different tissues of the body. The best studied set of isozymes is the lactate dehydrogenase (LDH) system. LDH is a tetrameric enzyme composed of all possible arrangements of two different protein subunits; the subunits are known as H (for heart) and M (for skeletal muscle). These subunits combine in various combinations leading to 5 distinct isozymes. The all H isozyme is characteristic of that from heart tissue, and the all M isozyme is typically found in skeletal muscle and liver. These isozymes all catalyze the same chemical reaction, but they exhibit differing degrees of efficiency. The detection of specific LDH isozymes in the blood is highly diagnostic of tissue damage such as occurs during cardiac infarct.
back to the top


--------------------------------------------------------------------------------

Enzyme-Substrate Interactions

The favored model of enzyme substrate interaction is known as the induced fit model. This model proposes that the initial interaction between enzyme and substrate is relatively weak, but that these weak interactions rapidly induce conformational changes in the enzyme that strengthen binding and bring catalytic sites close to substrate bonds to be altered. After binding takes place, one or more mechanisms of catalysis generates transition- state complexes and reaction products. The possible mechanisms of catalysis are four in number:



1. Catalysis by Bond Strain: In this form of catalysis, the induced structural rearrangements that take place with the binding of substrate and enzyme ultimately produce strained substrate bonds, which more easily attain the transition state. The new conformation often forces substrate atoms and bulky catalytic groups, such as aspartate and glutamate, into conformations that strain existing substrate bonds.



2. Catalysis by Proximity and Orientation: Enzyme-substrate interactions orient reactive groups and bring them into proximity with one another. In addition to inducing strain, groups such as aspartate are frequently chemically reactive as well, and their proximity and orientation toward the substrate thus favors their participation in catalysis.



3. Catalysis Involving Proton Donors (Acids) and Acceptors (Bases): Other mechanisms also contribute significantly to the completion of catalytic events initiated by a strain mechanism, for example, the use of glutamate as a general acid catalyst (proton donor).



4. Covalent Catalysis: In catalysis that takes place by covalent mechanisms, the substrate is oriented to active sites on the enzymes in such a way that a covalent intermediate forms between the enzyme or coenzyme and the substrate. One of the best-known examples of this mechanism is that involving proteolysis by serine proteases, which include both digestive enzymes (trypsin, chymotrypsin, and elastase) and several enzymes of the blood clotting cascade. These proteases contain an active site serine whose R group hydroxyl forms a covalent bond with a carbonyl carbon of a peptide bond, thereby causing hydrolysis of the peptide bond.
back to the top


--------------------------------------------------------------------------------

Chemical Reactions and Rates

According to the conventions of biochemistry, the rate of a chemical reaction is described by the number of molecules of reactant(s) that are converted into product(s) in a specified time period. Reaction rate is always dependent on the concentration of the chemicals involved in the process and on rate constants that are characteristic of the reaction. For example, the reaction in which A is converted to B is written as follows:

A ------> B

The rate of this reaction is expressed algebraically as either a decrease in the concentration of reactant A:



-[A] = k[B]

or an increase in the concentration of product B:

[B] = k[A]



In the second equation (of the 3 above) the negative sign signifies a decrease in concentration of A as the reaction progresses, brackets define concentration in molarity and the k is known as a rate constant. Rate constants are simply proportionality constants that provide a quantitative connection between chemical concentrations and reaction rates. Each chemical reaction has characteristic values for its rate constants; these in turn directly relate to the equilibrium constant for that reaction. Thus, reaction can be rewritten as an equilibrium expression in order to show the relationship between reaction rates, rate constants and the equilibrium constant for this simple case. The rate constant for the forward reaction is defined as k+1 and the reverse as k-1.

At equilibrium the rate (v) of the forward reaction (A -----> B) is--- by definition--- equal to that of the reverse or back reaction (B -----> A), a relationship which is algebraically symbolized as:

vforward = vreverse

where, for the forward reaction:

vforward = k+1[A]

and for the reverse reaction:

vreverse = k-1[B]



In the above equations, k+1 and k-1 represent rate constants for the forward and reverse reactions, respectively. The negative subscript refers only to a reverse reaction, not to an actual negative value for the constant. To put the relationships of the two equations into words, we state that the rate of the forward reaction [vforward] is equal to the product of the forward rate constant k+1 and the molar concentration of A. The rate of the reverse reaction is equal to the product of the reverse rate constant k-1 and the molar concentration of B.

At equilibrium, the rate of the forward reaction is equal to the rate of the reverse reaction leading to the equilibrium constant of the reaction and is expressed by:



[B]/[A] = k+1/k-1 = Keq

This equation demonstrates that the equilibrium constant for a chemical reaction is not only equal to the equilibrium ratio of product and reactant concentrations, but is also equal to the ratio of the characteristic rate constants of the reaction.
back to the top


--------------------------------------------------------------------------------

Chemical Reaction Order

Reaction order refers to the number of molecules involved in forming a reaction complex that is competent to proceed to product(s). Empirically, order is easily determined by summing the exponents of each concentration term in the rate equation for a reaction. A reaction characterized by the conversion of one molecule of A to one molecule of B with no influence from any other reactant or solvent is a first-order reaction. The exponent on the substrate concentration in the rate equation for this type of reaction is 1. A reaction with two substrates forming two products would a second-order reaction. However, the reactants in second- and higher- order reactions need not be different chemical species. An example of a second order reaction is the formation of ATP through the condensation of ADP with orthophosphate:



ADP + H2PO4 <----> ATP + H2O

For this reaction the forward reaction rate would be written as:

vforward = k1[ADP][H2PO4]

back to the top


--------------------------------------------------------------------------------

Enzymes as Biological Catalysts

In cells and organisms most reactions are catalyzed by enzymes, which are regenerated during the course of a reaction. These biological catalysts are physiologically important because they speed up the rates of reactions that would otherwise be too slow to support life. Enzymes increase reaction rates--- sometimes by as much as one millionfold, but more typically by about one thousand fold. Catalysts speed up the forward and reverse reactions proportionately so that, although the magnitude of the rate constants of the forward and reverse reactions is are increased, the ratio of the rate constants remains the same in the presence or absence of enzyme. Since the equilibrium constant is equal to a ratio of rate constants, it is apparent that enzymes and other catalysts have no effect on the equilibrium constant of the reactions they catalyze.

Enzymes increase reaction rates by decreasing the amount of energy required to form a complex of reactants that is competent to produce reaction products. This complex is known as the activated state or transition state complex for the reaction. Enzymes and other catalysts accelerate reactions by lowering the energy of the transition state. The free energy required to form an activated complex is much lower in the catalyzed reaction. The amount of energy required to achieve the transition state is lowered; consequently, at any instant a greater proportion of the molecules in the population can achieve the transition state. The result is that the reaction rate is increased.
back to the top


--------------------------------------------------------------------------------

Michaelis-Menton Kinetics

In typical enzyme-catalyzed reactions, reactant and product concentrations are usually hundreds or thousands of times greater than the enzyme concentration. Consequently, each enzyme molecule catalyzes the conversion to product of many reactant molecules. In biochemical reactions, reactants are commonly known as substrates. The catalytic event that converts substrate to product involves the formation of a transition state, and it occurs most easily at a specific binding site on the enzyme. This site, called the catalytic site of the enzyme, has been evolutionarily structured to provide specific, high-affinity binding of substrate(s) and to provide an environment that favors the catalytic events. The complex that forms, when substrate(s) and enzyme combine, is called the enzyme substrate (ES) complex. Reaction products arise when the ES complex breaks down releasing free enzyme.

Between the binding of substrate to enzyme, and the reappearance of free enzyme and product, a series of complex events must take place. At a minimum an ES complex must be formed; this complex must pass to the transition state (ES*); and the transition state complex must advance to an enzyme product complex (EP). The latter is finally competent to dissociate to product and free enzyme. The series of events can be shown thus:



E + S <---> ES <---> ES* <---> EP <---> E + P



The kinetics of simple reactions like that above were first characterized by biochemists Michaelis and Menten. The concepts underlying their analysis of enzyme kinetics continue to provide the cornerstone for understanding metabolism today, and for the development and clinical use of drugs aimed at selectively altering rate constants and interfering with the progress of disease states. The Michaelis-Menten equation is a quantitative description of the relationship among the rate of an enzyme- catalyzed reaction [v1], the concentration of substrate [S] and two constants, Vmax and Km (which are set by the particular equation). The symbols used in the Michaelis-Menton equation refer to the reaction rate [v1], maximum reaction rate (Vmax), substrate concentration [S] and the Michaelis-Menton constant (Km).

The Michaelis-Menten equation can be used to demonstrate that at the substrate concentration that produces exactly half of the maximum reaction rate, i.e.,1/2 Vmax, the substrate concentration is numerically equal to Km. This fact provides a simple yet powerful bioanalytical tool that has been used to characterize both normal and altered enzymes, such as those that produce the symptoms of genetic diseases. Rearranging the Michaelis-Menton equation leads to:




From this equation it should be apparent that when the substrate concentration is half that required to support the maximum rate of reaction, the observed rate, v1, will, be equal to Vmax divided by 2; in other words, v1 = [Vmax/2]. At this substrate concentration Vmax/v1 will be exactly equal to 2, with the result that:



[S](1) = Km



The latter is an algebraic statement of the fact that, for enzymes of the Michaelis-Menten type, when the observed reaction rate is half of the maximum possible reaction rate, the substrate concentration is numerically equal to the Michaelis-Menten constant. In this derivation, the units of Km are those used to specify the concentration of S, usually Molarity.

The Michaelis-Menten equation has the same form as the equation for a rectangular hyperbola; graphical analysis of reaction rate (v) versus substrate concentration [S] produces a hyperbolic rate plot.





Plot of substrate concentration versus reaction velocity




The key features of the plot are marked by points A, B and C. At high substrate concentrations the rate represented by point C the rate of the reaction is almost equal to Vmax, and the difference in rate at nearby concentrations of substrate is almost negligible. If the Michaelis-Menten plot is extrapolated to infinitely high substrate concentrations, the extrapolated rate is equal to Vmax. When the reaction rate becomes independent of substrate concentration, or nearly so, the rate is said to be zero order. (Note that the reaction is zero order only with respect to this substrate. If the reaction has two substrates, it may or may not be zero order with respect to the second substrate). The very small differences in reaction velocity at substrate concentrations around point C (near Vmax) reflect the fact that at these concentrations almost all of the enzyme molecules are bound to substrate and the rate is virtually independent of substrate, hence zero order. At lower substrate concentrations, such as at points A and B, the lower reaction velocities indicate that at any moment only a portion of the enzyme molecules are bound to the substrate. In fact, at the substrate concentration denoted by point B, exactly half the enzyme molecules are in an ES complex at any instant and the rate is exactly one half of Vmax. At substrate concentrations near point A the rate appears to be directly proportional to substrate concentration, and the reaction rate is said to be first order.
back to the top


--------------------------------------------------------------------------------

Inhibition of Enzyme Catalyzed Reactions

To avoid dealing with curvilinear plots of enzyme catalyzed reactions, biochemists Lineweaver and Burk introduced an analysis of enzyme kinetics based on the following rearrangement of the Michaelis-Menten equation:

[1/v] = [Km (1)/ Vmax[S] + (1)/Vmax]



Plots of 1/v versus 1/[S] yield straight lines having a slope of Km/Vmax and an intercept on the ordinate at 1/Vmax.







A Lineweaver-Burk Plot




An alternative linear transformation of the Michaelis-Menten equation is the Eadie-Hofstee transformation:



v/[S] = -v [1/Km] + [Vmax/Km]

and when v/[S] is plotted on the y-axis versus v on the x-axis, the result is a linear plot with a slope of -1/Km and the value Vmax/Km as the intercept on the y-axis and Vmax as the intercept on the x-axis.

Both the Lineweaver-Burk and Eadie-Hofstee transformation of the Michaelis-Menton equation are useful in the analysis of enzyme inhibition. Since most clinical drug therapy is based on inhibiting the activity of enzymes, analysis of enzyme reactions using the tools described above has been fundamental to the modern design of pharmaceuticals. Well- known examples of such therapy include the use of methotrexate in cancer chemotherapy to semi-selectively inhibit DNA synthesis of malignant cells, the use of aspirin to inhibit the synthesis of prostaglandins which are at least partly responsible for the aches and pains of arthritis, and the use of sulfa drugs to inhibit the folic acid synthesis that is essential for the metabolism and growth of disease-causing bacteria. In addition, many poisons, such as cyanide, carbon monoxide and polychlorinated biphenols (PCBs). produce their life- threatening effects by means of enzyme inhibition.

Enzyme inhibitors fall into two broad classes: those causing irreversible inactivation of enzymes and those whose inhibitory effects can be reversed. Inhibitors of the first class usually cause an inactivating, covalent modification of enzyme structure. Cyanide is a classic example of an irreversible enzyme inhibitor: by covalently binding mitochondrial cytochrome oxidase, it inhibits all the reactions associated with electron transport. The kinetic effect of irreversible inhibitors is to decrease the concentration of active enzyme, thus decreasing the maximum possible concentration of ES complex. Since the limiting enzyme reaction rate is often k2[ES], it is clear that under these circumstances the reduction of enzyme concentration will lead to decreased reaction rates. Note that when enzymes in cells are only partially inhibited by irreversible inhibitors, the remaining unmodified enzyme molecules are not distinguishable from those in untreated cells; in particular, they have the same turnover number and the same Km. Turnover number, related to Vmax, is defined as the maximum number of moles of substrate that can be converted to product per mole of catalytic site per second. Irreversible inhibitors are usually considered to be poisons and are generally unsuitable for therapeutic purposes.

Reversible inhibitors can be divided into two main categories; competitive inhibitors and noncompetitive inhibitors, with a third category, uncompetitive inhibitors, rarely encountered.



Inhibitor Type
Binding Site on Enzyme
Kinetic effect

Competitive Inhibitor
Specifically at the catalytic site, where it competes with substrate for binding in a dynamic equilibrium- like process. Inhibition is reversible by substrate.
Vmax is unchanged; Km, as defined by [S] required for 1/2 maximal activity, is increased.

Noncompetitive Inhibitor
Binds E or ES complex other than at the catalytic site. Substrate binding unaltered, but ESI complex cannot form products. Inhibition cannot be reversed by substrate.
Km appears unaltered; Vmax is decreased proportionately to inhibitor concentration.

Uncompetitive Inhibitor
Binds only to ES complexes at locations other than the catalytic site. Substrate binding modifies enzyme structure, making inhibitor- binding site available. Inhibition cannot be reversed by substrate.
Apparent Vmax decreased; Km, as defined by [S] required for 1/2 maximal activity, is decreased.




The hallmark of all the reversible inhibitors is that when the inhibitor concentration drops, enzyme activity is regenerated. Usually these inhibitors bind to enzymes by non-covalent forces and the inhibitor maintains a reversible equilibrium with the enzyme. The equilibrium constant for the dissociation of enzyme inhibitor complexes is known as KI:



KI = [E][I]/[E--I--complex]



The importance of KI is that in all enzyme reactions where substrate, inhibitor and enzyme interact, the normal Km and or Vmax for substrate enzyme interaction appear to be altered. These changes are a consequence of the influence of KI on the overall rate equation for the reaction. The effects of KI are best observed in Lineweaver-Burk plots.

Probably the best known reversible inhibitors are competitive inhibitors, which always bind at the catalytic or active site of the enzyme. Most drugs that alter enzyme activity are of this type. Competitive inhibitors are especially attractive as clinical modulators of enzyme activity because they offer two routes for the reversal of enzyme inhibition, while other reversible inhibitors offer only one. First, as with all kinds of reversible inhibitors, a decreasing concentration of the inhibitor reverses the equilibrium regenerating active free enzyme. Second, since substrate and competitive inhibitors both bind at the same site they compete with one another for binding

Raising the concentration of substrate (S), while holding the concentration of inhibitor constant, provides the second route for reversal of competitive inhibition. The greater the proportion of substrate, the greater the proportion of enzyme present in competent ES complexes.

As noted earlier, high concentrations of substrate can displace virtually all competitive inhibitor bound to active sites. Thus, it is apparent that Vmax should be unchanged by competitive inhibitors. This characteristic of competitive inhibitors is reflected in the identical vertical-axis intercepts of Lineweaver-Burk plots, with and without inhibitor.





Lineweaver-Burk Plots of Inhibited Enzymes




Since attaining Vmax requires appreciably higher substrate concentrations in the presence of competitive inhibitor, Km (the substrate concentration at half maximal velocity) is also higher, as demonstrated by the differing negative intercepts on the horizontal axis in panel B.

Analogously, panel C illustrates that noncompetitive inhibitors appear to have no effect on the intercept at the x-axis implying that noncompetitive inhibitors have no effect on the Km of the enzymes they inhibit. Since noncompetitive inhibitors do not interfere in the equilibration of enzyme, substrate and ES complexes, the Km's of Michaelis-Menten type enzymes are not expected to be affected by noncompetitive inhibitors, as demonstrated by x-axis intercepts in panel C. However, because complexes that contain inhibitor (ESI) are incapable of progressing to reaction products, the effect of a noncompetitive inhibitor is to reduce the concentration of ES complexes that can advance to product. Since Vmax = k2[Etotal], and the concentration of competent Etotal is diminished by the amount of ESI formed, noncompetitive inhibitors are expected to decrease Vmax, as illustrated by the y-axis intercepts in panel C.

A corresponding analysis of uncompetitive inhibition leads to the expectation that these inhibitors should change the apparent values of Km as well as Vmax. Changing both constants leads to double reciprocal plots, in which intercepts on the x and y axes are proportionately changed; this leads to the production of parallel lines in inhibited and uninhibited reactions.
back to the top


--------------------------------------------------------------------------------

Regulation of Enzyme Activity

While it is clear that enzymes are responsible for the catalysis of almost all biochemical reactions, it is important to also recognize that rarely, if ever, do enzymatic reactions proceed in isolation. The most common scenario is that enzymes catalyze individual steps of multi-step metabolic pathways, as is the case with glycolysis, gluconeogenesis or the synthesis of fatty acids. As a consequence of these lock- step sequences of reactions, any given enzyme is dependent on the activity of preceding reaction steps for its substrate.

In humans, substrate concentration is dependent on food supply and is not usually a physiologically important mechanism for the routine regulation of enzyme activity. Enzyme concentration, by contrast, is continually modulated in response to physiological needs. Three principal mechanisms are known to regulate the concentration of active enzyme in tissues:

· 1. Regulation of gene expression controls the quantity and rate of enzyme synthesis.

· 2. Proteolytic enzyme activity determines the rate of enzyme degradation.

· 3. Covalent modification of preexisting pools of inactive proenzymes produces active enzymes.

Enzyme synthesis and proteolytic degradation are comparatively slow mechanisms for regulating enzyme concentration, with response times of hours, days or even weeks. Proenzyme activation is a more rapid method of increasing enzyme activity but, as a regulatory mechanism, it has the disadvantage of not being a reversible process. Proenzymes are generally synthesized in abundance, stored in secretory granules and covalently activated upon release from their storage sites. Examples of important proenzymes include pepsinogen, trypsinogen and chymotrypsinogen, which give rise to the proteolytic digestive enzymes. Likewise, many of the proteins involved in the cascade of chemical reactions responsible for blood clotting are synthesized as proenzymes. Other important proteins, such as peptide hormones and collagen, are also derived by covalent modification of precursors.

Another mechanism of regulating enzyme activity is to sequester enzymes in compartments where access to their substrates is limited. For example, the proteolysis of cell proteins and glycolipids by enzymes responsible for their degradation is controlled by sequestering these enzymes within the lysosome.

In contrast to regulatory mechanisms that alter enzyme concentration, there is an important group of regulatory mechanisms that do not affect enzyme concentration, are reversible and rapid in action, and actually carry out most of the moment- to- moment physiological regulation of enzyme activity. These mechanisms include allosteric regulation, regulation by reversible covalent modification and regulation by control proteins such as calmodulin. Reversible covalent modification is a major mechanism for the rapid and transient regulation of enzyme activity. The best examples, again, come from studies on the regulation of glycogen metabolism where phosphorylation of glycogen synthase and glycogen phosphorylase kinase results in the stimulation of glycogen degradation while glycogen synthesis is coordinately inhibited. Numerous other enzymes of intermediary metabolism are affected by phosphorylation, either positively or negatively. These covalent phosphorylations can be reversed by a separate sub-subclass of enzymes known as phosphatases. Recent research has indicated that the aberrant phosphorylation of growth factor and hormone receptors, as well as of proteins that regulate cell division, often leads to unregulated cell growth or cancer. The usual sites for phosphate addition to proteins are the serine, threonine and tyrosine R group hydroxyl residues.
back to the top


--------------------------------------------------------------------------------

Allosteric Enzymes

In addition to simple enzymes that interact only with substrates and inhibitors, there is a class of enzymes that bind small, physiologically important molecules and modulate activity in ways other than those described above. These are known as allosteric enzymes; the small regulatory molecules to which they bind are known as effectors. Allosteric effectors bring about catalytic modification by binding to the enzyme at distinct allosteric sites, well removed from the catalytic site, and causing conformational changes that are transmitted through the bulk of the protein to the catalytically active site(s).

The hallmark of effectors is that when they bind to enzymes, they alter the catalytic properties of an enzyme's active site. Those that increase catalytic activity are known as positive effectors. Effectors that reduce or inhibit catalytic activity are negative effectors.

Most allosteric enzymes are oligomeric (consisting of multiple subunits); generally they are located at or near branch points in metabolic pathways, where they are influential in directing substrates along one or another of the available metabolic paths. The effectors that modulate the activity of these allosteric enzymes are of two types. Those activating and inhibiting effectors that bind at allosteric sites are called heterotropic effectors. (Thus there exist both positive and negative heterotropic effectors.) These effectors can assume a vast diversity of chemical forms, ranging from simple inorganic molecules to complex nucleotides such as cyclic adenosine monophosphate (cAMP). Their single defining feature is that they are not identical to the substrate.

In many cases the substrate itself induces distant allosteric effects when it binds to the catalytic site. Substrates acting as effectors are said to be homotropic effectors. When the substrate is the effector, it can act as such, either by binding to the substrate-binding site, or to an allosteric effector site. When the substrate binds to the catalytic site it transmits an activity-modulating effect to other subunits of the molecule. Often used as the model of a homotropic effector is hemoglobin, although it is not a branch-point enzyme and thus does not fit the definition on all counts.

There are two ways that enzymatic activity can be altered by effectors: the Vmax can be increased or decreased, or the Km can be raised or lowered. Enzymes whose Km is altered by effectors are said to be K-type enzymes and the effector a K-type effector. If Vmax is altered, the enzyme and effector are said to be V-type. Many allosteric enzymes respond to multiple effectors with V-type and K-type behavior. Here again, hemoglobin is often used as a model to study allosteric interactions, although it is not strictly an enzyme.

In the preceding discussion we assumed that allosteric sites and catalytic sites were homogeneously present on every subunit of an allosteric enzyme. While this is often the case, there is another class of allosteric enzymes that are comprised of separate catalytic and regulatory subunits. The archetype of this class of enzymes is cAMP-dependent protein kinase (PKA), whose mechanism of activation is illustrated in the Figure below. The enzyme is tetrameric, containing two catalytic subunits and two regulatory subunits, and enzymatically inactive. When intracellular cAMP levels rise, one molecule of cAMP binds to each regulatory subunit, causing the tetramer to dissociate into one regulatory dimer and two catalytic monomers. In the dissociated form, the catalytic subunits are fully active; they catalyze the phosphorylation of a number of other enzymes, such as those involved in regulating glycogen metabolism. The regulatory subunits have no catalytic activity.