Amino Acids

  • May 2020
  • PDF

This document was uploaded by user and they confirmed that they have the permission to share it. If you are author or own the copyright of this book, please report to us by using this DMCA report form. Report DMCA


Overview

Download & View Amino Acids as PDF for free.

More details

  • Words: 3,142
  • Pages: 11
In chemistry, an amino acid is a molecule containing both amine and carboxyl functional groups. These molecules are particularly important in biochemistry, where this term refers to alpha-amino acids with the general formula H2NCHRCOOH, where R is an organic substituent. In the alpha amino acids, the amino and carboxylate groups are attached to the same carbon, which is called the α–carbon.The various alpha amino acids differ in which side chain (R group) is attached to their alpha carbon. They can vary in size from just a hydrogen atom in glycine through a methyl group in alanine to a large heterocyclic group in tryptophan. Amino acids are critical to life, and have a variety of roles I metabolism. One particularly important function is as the building blocks of proteins, which are linear chains of amino acids. Amino acids are also important in many other biological molecules, such as forming parts of coenzymes, as in S-adenosylmethionine, or as precursors for the biosynthesis of molecules such as heme. Due to this central role in biochemistry, amino acids are very important in nutrition. The amino acids are commonly used in food technology and industry. For example, monosodium glutamate is a common flavor enhancer that gives foods the taste called umami. Beyond the amino acids that are found in all forms of life, amino acids are also used in industry, with the production of biodegradable plastics, drugs and chiral catalysts being particularly important applications.

Occurrence and functions in biochemistry

A polypeptide is a chain of amino acids.

Standard amino acids See also: Primary structure and Posttranslational modification Amino acids are the basic structural building units of proteins. They form short polymer chains called peptides or longer chains called either polypeptides or proteins. These polymers are linear and unbranched. The process of making proteins is called translation

and involves the step-by-step addition of amino acids to a growing protein chain by a ribozyme that is called a ribosome.[13] The order in which the amino acids are added is read through the genetic code from an mRNA template, which is a RNA copy of one of the organism's genes. Twenty amino acids are encoded by the standard genetic code and are called proteinogenic or standard amino acids.[3]

The amino acid selenocysteine

Non-standard amino acids Aside from the twenty standard amino acids, there are a vast number of "non-standard" amino acids. Two of these can be specified by the genetic code, but are rather rare in proteins. Selenocysteine is incorporated into some proteins at a UGA codon, which is normally a stop codon.[14] Pyrrolysine is used by some methanogenic archaea in enzymes that they use to produce methane. It is coded for with the codon UAG.[15] Other nonstandard amino acids found in proteins are formed by post-translational modification, which is modification after translation in protein synthesis. These modifications are often essential for the function or regulation of a protein; for example, the carboxylation of glutamate allows for better binding of calcium cations,[16] and the hydroxylation of proline is critical for maintaining connective tissues.[17] Another example is the formation of hypusine in the translation initiation factor EIF5A, through modification of a lysine residue.[18] Such modifications can also determine the localization of the protein, e.g., the addition of long hydrophobic groups can cause a protein to bind to a phospholipid membrane.[19]

β-alanine and its α-alanine isomer. Examples of nonstandard amino acids that are not found in proteins include lanthionine, 2-aminoisobutyric acid, dehydroalanine and the neurotransmitter gamma-aminobutyric acid. Nonstandard amino acids often occur as intermediates in the metabolic pathways for

standard amino acids — for example ornithine and citrulline occur in the urea cycle, part of amino acid catabolism (see below).[20] A rare exception to the dominance of α-amino acids in biology is the β-amino acid beta alanine (3-aminopropanoic acid), which is used in plants and microorganisms in the synthesis of pantothenic acid (vitamin B5), a component of coenzyme A.[21]

In human nutrition Further information: Protein in nutrition and Amino acid synthesis When taken up into the body in the diet, the 20 standard amino acids are either used to synthesize proteins and other biomolecules or oxidized to urea and carbon dioxide as a source of energy.[22] The oxidation pathway starts with the removal of the amino group by a transaminase, the amino group is then fed into the urea cycle. The other product of transamidation is a keto acid that enters the citric acid cycle.[23] Glucogenic amino acids can also be converted into glucose, through gluconeogenesis.[24] Of the 20 standard amino acids, 8 are called essential amino acids because the human body cannot synthesize them from other compounds at the level needed for normal growth, so they must be obtained from food.[25] However, the situation is quite complicated since cysteine, taurine, tyrosine, histidine and arginine are semiessential amino acids in children, because the metabolic pathways that synthesize these amino acids are not fully developed.[26][27] The amounts required also depend on the age and health of the individual, so it is hard to make general statements about the dietary requirement for some amino acids. Essential Isoleucine Leucine Lysine Methionine Phenylalanine Threonine Tryptophan Valine

Nonessential Alanine Asparagine Aspartate Cysteine* Glutamate Glutamine* Glycine* Proline* Serine* Tyrosine* Arginine* Histidine*

(*) Essential only in certain cases.[28][29] Several common mnemonics have evolved for remembering the amino acids often described as essential in various contexts. PVT TIM HALL ("Private Tim Hall") uses the first letter of each of these amino acids.[30] Another mnemonic that frequently occurs in

student practice materials beneath "AH TV TILL Past Midnight", is "These ten valuable amino acids have long preserved life in man".[31]

Non-protein functions Further information: Amino acid neurotransmitter In humans, non-protein amino acids also have important roles as metabolic intermediates, such as in the biosynthesis of the neurotransmitter gamma-aminobutyric acid. Many amino acids are used to synthesize other molecules, for example: • • • • •

Tryptophan is a precursor of the neurotransmitter serotonin.[32] Glycine is a precursor of porphyrins such as heme.[33] Arginine is a precursor of nitric oxide.[34] Ornithine and S-adenosylmethionine are precursors of polyamines.[35] Aspartate, glycine and glutamine are precursors of nucleotides.[36]

However, not all of the functions of other abundant non-standard amino acids are known, for example taurine is a major amino acid in muscle and brain tissues, but although many functions have been proposed, its precise role in the body has not been determined.[37] Some non-standard amino acids are used as defenses against herbivores in plants.[38] For example canavanine is an analogue of arginine that is found in many legumes,[39] and in particularly large amounts in Canavalia gladiata (sword bean).[40] This amino acid protects the plants from predators such as insects and can cause illness in people if some types of legumes are eaten without processing.[41] The non-protein amino acid mimosine is found in other species of legume, particularly Leucaena leucocephala.[42] This compound is an analogue of tyrosine and can poison animals that graze on these plants.

Amino Acids and Proteins Functions of Proteins Because proteins are a highly evolved and diverse class of molecules, they perform endless tasks and functions within both plants and animals. They are important in the biosyntheses of hormones, enzymes, and membrane channels and pumps. In animals, proteins also function in the immune system and can be used in the production of energy. In essence, proteins are the currency of life.

Biosyntheses: essential and nonessential amino acids (transanimation)

Since proteins constitute the majority of tissues in the body and since these tissues are constantly in protein flux, proteins are degraded and synthesized within all tissues on a regular basis. Some of the amino acids that are degraded can be recycled by the liver and used again for other biosyntheses, but a significant portion of this protein cannot be replaced. Through a process known as transamination, the liver synthesizes amino acids.

Figure 2.1: Transamination During this reaction, an amino group from glutamic acid is transferred to an alpha keto acid, which is a precursor for amino acid sythesis. Aminotransferases, which are derived from vitamin B6, are the enzyme responsible for the reaction. The amino acids that can be produced through transanimation include alanine, arginine, asparagine, aspartic acid, cysteine, glutamic acid, glutamine, glycine, proline, serine, and tyrosine. These are obviously the non-essential amino acids, since they can be synthesized in the body.

Energy: ketogenic and glucogenic

When the body's energy sources are low, it begins to degrade proteins for use as an alternative energy source. Amino acids can be classified as glucogenic or ketogenic.

Glucogenic Amino Acids Glucogenic amino acids can be degraded to pyruvate or an intermediate in the Krebs Cycle. They are named glucogenic because they can produce glucose under conditions of low glucose. This process is also known as gluconeogenesis, or the production of "new glucose." Amino acids form glucose through degradation to pyruvate or an intermediate in the Krebs Cycle.

Figure 2.2: Amino acid degradation to pyruvate The intermediates can then be converted to oxaloacetate, the main precursor for gluconeogenesis. The following amino acids are glucogenic: alanine, cysteine, glycine, serine, threonine, tryptophan, asparagine, aspartate, phenylalanine, tyrosine, isoleucine, methionine, threonine, valine, arginine, glutamate, glutamine, histidine, and proline.

Ketogenic Amino Acids

In contrast, ketogenic amino acids can produce ketones when energy sources are low. Some of these amino acids are degraded directly to ketone bodies such as acetoacetate (see ). They include leucine, lysine, phenylalanine, tryptophan, and tyrosine. The other ketogenic amino acids can be converted to acetyl CoA. Acetyl CoA has several different fates, one of which is the conversion to acetoacetate. Although not a preferential energy source, acetoacetate can be metabolized by the brain and muscle for energy when blood glucose is low. Acetoacetate cannot be used in gluconeogenesis, since acetyl CoA cannot be converted directly to oxaloacetate. When energy sources are high, both glucogenic and ketogenic amino acids are converted to fatty acids through the intermediate acetyl CoA. Other amino acids that are degraded to intermediates in the Krebs Cycle are siphoned off into the production of urea, a nitrogenous carboxyl compound that is filtered through the kidneys and secreted in the urine.

Maintaining fluid balance Blood proteins such as albumin and globulin function to maintain fluid balance in the body. When the concentrations of proteins in the bloodstream are low, the fluid in the blood (serum) begins to seep into surrounding tissue. Proteins in the blood can counteract this effect by increasing the osmotic potential and forcing fluid back into the bloodstream. Therefore, low amounts of protein in the blood cause edema, a condition that is characterized by an abnormal amount of fluid in the tissue and extracellular space. Edema is seen in starvation, low calorie diets, and diseases like AIDS that decrease the amount of circulating antibodies and albumin.

Hormones Many hormones are composed of polypeptide chains. The beta cells in the pancreas produce the peptide hormone insulin. Insulin facilitates glucose uptake in cells and promotes the synthesis of glycogen and fatty acids. Diabetics must inject the peptide hormone insulin because it will be degraded into amino acids in the stomach and small intestine if taken orally. Other examples of peptide signal molecules include neurotransmitters, a class of molecules that are produced and released at nerve endings in the brain and autonomic nervous system.

Enzymes

Enzymes are an entirely different class of proteins. Enzymes catalyze biological reactions by increasing the reaction rates by factors of at least a million. Since most reactions in the body proceed at imperceptible rates without enzymes, it is critical that these proteins be present in sufficient quantities for proper functioning of cells. How are enzymes so efficient and fast? First, enzymes are highly specific for their substrates. Substrates are any molecules for which the enzyme binds preferentially or with a high affinity. Enzymes are able to bind specific substrates because they form deep pockets or clefts that are complementary to the three dimensional substrate conformations. Since the enzyme pocket is complementary to the substrate, a large number of noncovalent, dipole-dipole, and Van der Waal interactions can occur, favoring enzyme-substrate binding. The second reason for the high catalytic rate of enzymes is that they are able to stabilize transition state intermediates. By stabilizing these intermediates, enzymes are able to decrease the activation energy required for the reaction to occur. Upon reaching its highenergy transitional state, the bound substrate can be easily converted to the cell's desired product, which is then released to meet the needs of the cell. These reactions can take place on the order of microseconds to nanoseconds. In fact, many enzymes are so efficient and fast that they approach the diffusion-controlled limit, the rate at which substrate diffusion cannot keep up with the rate at which the enzyme catalyzes the reaction. Enzymes like these have reached catalytic perfection. Besides their active sites, many enzymes have other locations or crevices where molecules can bind. Regulatory sites, also called allosteric sites, are places other than the active site of the enzyme that serve to regulate enzymatic activity.

Allosteric sites as inhibitors The end product in a series of reactions catalyzed by different enzymes at each step may bind to the allosteric site and inhibit the activity of the first enzyme in the pathway. When an inhibitory molecule binds to an allosteric region on the enzyme it can cause the active site of the enzyme to close up and become inactive.

Figure 2.3: Enzymatic negative feedback loop

This type of negative feedback is used to control levels of products from reaching levels that are too high or unnecessary.

Allosteric sites as stimulators Allosteric sites can also be areas that cause the stimulation of enzymatic rates. When these allosteric sites are occupied, the enzyme's active site may change in shape, becoming more efficient or receptive as a catalyst.

Covalent Modification Further regulation of enzymes comes in the form of covalent modification. Many enzymes are regulated by reversible attachment of phosphoryl groups to serine andthreonine amino acid residues. Specific types of enzymes called kinases phosphorylate add phosphoryl groups to other enzymes while phosphatases remove phosphate groups. By adding just one covalent bond to the enzyme, its activity can be drastically changed. For example, during levels of low blood glucose, glucagon and epinephrine are secreted into the blood and bind to muscle and brain cell receptors. Upon binding, these hormones cause a cascade of effects within the cell, which results in the phosphorylation of a number of proteins, including a series of enzymes involved in metabolism. All phosphorylations act to increase the rate of enzymes involved in glycogen and triglyceride degradation while inhibiting the rate of enzymes involved in glycolysis and the Citric Acid Cycle. In effect, the addition of phosphate groups to these enzymes causes the level of blood glucose to increase.

Membrane channels and pumps Proteins are also abundant within biological membranes. Many cellular receptors, channels, and pumps are bound to membranes. Since these proteins span across a nonpolar environment, many of their residues facing this environment are also nonpolar, allowing more favorable interactions to occur. Both channels and pumps are involved in the regulation of fluids and ions within and outside the cell. However, they differ in many key respects. Channels allow ions to flow from an area of high concentration to an area of low concentration. This is a completely passive process. Pumps, on the other hand, force ions up their concentration gradient from a region of lower concentration to a region of higher concentration. This process is called active transport and usually requires the energy of adenosine triphosphate (ATP) in order to overcome the energy barrier. A classic example of a pump is the sodium potassium pump. Since during the excitation of neurons sodium ions are constantly traversing into the cell and potassium ions are

leaving the cell, the resting levels of these ions must be constantly restored. Unlike potassium, which is present in larger proportions outside the cell, sodium is in much higher concentrations outside the cell. The sodium potassium pump pushes sodium and potassium against their concentration gradients by binding ATP and hyrolyzing it for energy.

Immune function Proteins are also important in the immune response. There are two types of important proteins the immune system uses to scan the vast network of molecules in our body and determine self from nonself. In the first stage of the immune response, the body recognizes circulating foreign particles (antigens) through antibodies, which are produced by plasma cells (B lymphocytes). The ability of plasma cells to produce millions of different antibodies is inherent from birth; virtually any circulating antigen will be bound by its complementary antibody. Once bound by antibody, the antigen can be either consumed by macrophages or be further bound by mature plasma cells to stimulate the production of even more antibodies. Mature cells are stimulated to divide and form clones, which act as an immunological memory against further attack from identical antigens. Like enzymes, the amino acid sequence of the antibody determines its specificity. The whole recognition and response of plasma cells to a new infection is called the humoral response. The second stage of immune response is called the cellular immune response. In the cellular response, T lymphocytes (killer T cells) bind to foreign particles displayed on the surface of cells and destroy the contaminated cell. Helper T cells also bind to the foreign particles displayed on the surface of cells and stimulate the humoral response by helping plasma cells to proliferate. Why is the cellular response necessary if antibodies can recognize antigens and mark them for destruction? The answer lies in the fact that many of the viruses and bacteria that invade the body are found in greater concentrations within cells, preventing antibodies from reaching them. The immune system has adapted to this problem by cutting some of the foreign particles up into peptides to be displayed at the surface of the infected cell by a protein known as the major histocompatibility complex (MHC). Killer and helper T cells are specialized for recognizing peptides bound to these proteins, increasing the speed and effectiveness of the immune response.

SOURCE Dietary sources of protein include meats, eggs, nuts, grains, legumes, and dairy products such as milk and cheese.[1] Of the 20 amino acids used by humans in protein synthesis, 11 "nonessential" amino acids can be synthesized in sufficient quantities by the adult body, and are not required in the diet (though there are exceptions for some in special cases). The nine essential amino acids[2], plus arginine for the young[3], cannot be created by the body and must come from dietary sources.

Most animal sources and certain vegetable sources have the complete complement of all the essential amino acids in adequate proportions. However, it is not necessary to consume a single food source that contains all the essential amino acids, as long as all the essential amino acids are eventually present in the diet: see complete protein and protein combining.

Related Documents

Amino Acids
May 2020 26
Amino Acids
November 2019 33
Amino Acids
November 2019 27
Amino Acids
May 2020 15
Amino Acids
December 2019 30
The Amino Acids
December 2019 32