Amino acid

Amino acid
The generic structure of an alpha amino acid in its unionized form
Table of Amino Acids.
The 21 amino acids found in eukaryotes, grouped according to their side-chains' pKas and charge at physiological pH 7.4

Amino acids (pronounced /əˈmiːnoʊ ..., əˈmaɪnoʊ ..., ˈæmɪnoʊ .../) are molecules containing an amine group, a carboxylic acid group and a side-chain that varies between different amino acids. The key elements of an amino acid are carbon, hydrogen, oxygen, and nitrogen. They are particularly important in biochemistry, where the term usually refers to alpha-amino acids.

An alpha-amino acid has the generic formula H2NCHRCOOH, where R is an organic substituent;[1] the amino group is attached to the carbon atom immediately adjacent to the carboxylate group (the α–carbon). Other types of amino acid exist when the amino group is attached to a different carbon atom; for example, in gamma-amino acids (such as gamma-amino-butyric acid) the carbon atom to which the amino group attaches is separated from the carboxylate group by two other carbon atoms. The various alpha-amino acids differ in which side-chain (R-group) is attached to their alpha carbon, and can vary in size from just one hydrogen atom in glycine to a large heterocyclic group in tryptophan.

Amino acids are critical to life, and have many functions in metabolism. One particularly important function is to serve as the building blocks of proteins, which are linear chains of amino acids. Amino acids can be linked together in varying sequences to form a vast variety of proteins.[2] Twenty-two amino acids are naturally incorporated into polypeptides and are called proteinogenic or standard amino acids. Of these, 20 are encoded by the universal genetic code. Nine standard amino acids are called "essential" for humans because they cannot be created from other compounds by the human body, and so must be taken in as food.

Due to their central role in biochemistry, amino acids are important in nutrition and are commonly used in nutrition supplements, fertilizers, food technology and industry. In industry, applications include the production of biodegradable plastics, drugs, and chiral catalysts.

Contents

History

The first few amino acids were discovered in the early 19th century. In 1806, the French chemists Louis-Nicolas Vauquelin and Pierre Jean Robiquet isolated a compound in asparagus that proved to be asparagine, the first amino acid to be discovered.[3][4] Another amino acid that was discovered in the early 19th century was cystine, in 1810,[5] although its monomer, cysteine, was discovered much later, in 1884.[4][6] Glycine and leucine were also discovered around this time, in 1820.[7] Usage of the term amino acid in the English language is from 1898.[8]

General structure

Lysine contains six carbon atoms. The central carbon atom connected to the amino and carboxyl groups is labeled alpha. The four carbon atoms in its linear side-chain are labeled from beta (closest to the central carbon), gamma, delta, through to the epsilon carbon at the end of the chain and furthest from the central carbon.
Lysine with the carbon atoms in the side-chain labeled

In the structure shown at the top of the page, R represents a side-chain specific to each amino acid. The carbon atom next to the carboxyl group is called the α–carbon and amino acids with a side-chain bonded to this carbon are referred to as alpha amino acids. These are the most common form found in nature. In the alpha amino acids, the α–carbon is a chiral carbon atom, with the exception of glycine.[9] In amino acids that have a carbon chain attached to the α–carbon (such as lysine, shown to the right) the carbons are labeled in order as α, β, γ, δ, and so on.[10] In some amino acids, the amine group is attached to the β or γ-carbon, and these are therefore referred to as beta or gamma amino acids.

Amino acids are usually classified by the properties of their side-chain into four groups. The side-chain can make an amino acid a weak acid or a weak base, and a hydrophile if the side-chain is polar or a hydrophobe if it is nonpolar.[9] The chemical structures of the 22 standard amino acids, along with their chemical properties, are described more fully in the article on these proteinogenic amino acids.

The phrase "branched-chain amino acids" or BCAA refers to the amino acids having aliphatic side-chains that are non-linear; these are leucine, isoleucine, and valine. Proline is the only proteinogenic amino acid whose side-group links to the α-amino group and, thus, is also the only proteinogenic amino acid containing a secondary amine at this position.[9] In chemical terms, proline is, therefore, an imino acid, since it lacks a primary amino group,[11] although it is still classed as an amino acid in the current biochemical nomenclature,[12] and may also be called an "N-alkylated alpha-amino acid".[13]

Animation of two mirror image molecules rotating around a central axis.
The two optical isomers of alanine, D-Alanine and L-Alanine

Isomerism

Of the standard α-amino acids, all but glycine can exist in either of two optical isomers, called L or D amino acids, which are mirror images of each other (see also Chirality). While L-amino acids represent all of the amino acids found in proteins during translation in the ribosome, D-amino acids are found in some proteins produced by enzyme posttranslational modifications after translation and translocation to the endoplasmic reticulum, as in exotic sea-dwelling organisms such as cone snails.[14] They are also abundant components of the peptidoglycan cell walls of bacteria,[15] and D-serine may act as a neurotransmitter in the brain.[16] The L and D convention for amino acid configuration refers not to the optical activity of the amino acid itself, but rather to the optical activity of the isomer of glyceraldehyde from which that amino acid can, in theory, be synthesized (D-glyceraldehyde is dextrorotary; L-glyceraldehyde is levorotary). In alternative fashion, the (S) and (R) designators are used to indicate the absolute stereochemistry. Almost all of the amino acids in proteins are (S) at the α carbon, with cysteine being (R) and glycine non-chiral.[17] Cysteine is unusual since it has a sulfur atom at the second position in its side-chain, which has a larger atomic mass than the groups attached to the first carbon, which is attached to the α-carbon in the other standard amino acids, thus the (R) instead of (S).

An amino acid, which shown in two ionization states. First, it is shown in the same arrangement as the lead image. This is the unionised form. It is also shown in the ionized form, after the carboxyl group has lost a hydrogen atom, which introduces a negative charge, and the amino group has gained a hydrogen, which introduces a positive charge.
An amino acid in its (1) unionized and (2) zwitterionic forms

Zwitterions

The amine and carboxylic acid functional groups found in amino acids allow them to have amphiprotic properties.[9] Carboxylic acid groups (-CO2H) can be deprotonated to become negative carboxylates (-CO2- ), and α-amino groups (NH2-) can be protonated to become positive α-ammonium groups (+NH3-). At pH values greater than the pKa of the carboxylic acid group (mean for the 20 common amino acids is about 2.2, see the table of amino acid structures above), the negative carboxylate ion predominates. At pH values lower than the pKa of the α-ammonium group (mean for the 20 common α-amino acids is about 9.4), the nitrogen is predominantly protonated as a positively charged α-ammonium group. Thus, at pH between 2.2 and 9.4, the predominant form adopted by α-amino acids contains a negative carboxylate and a positive α-ammonium group, as shown in structure (2) on the right, so has net zero charge. This molecular state is known as a zwitterion, from the German Zwitter meaning hermaphrodite or hybrid.[18] Below pH 2.2, the predominant form will have a neutral carboxylic acid group and a positive α-ammonium ion (net charge +1), and above pH 9.4, a negative carboxylate and neutral α-amino group (net charge -1). The fully neutral form (structure (1) on the right) is a very minor species in aqueous solution throughout the pH range (less than 1 part in 107). Amino acids also exist as zwitterions in the solid phase, and crystallize with salt-like properties unlike typical organic acids or amines.

Isoelectric point

At pH values between the two pKa values, the zwitterion predominates, but coexists in dynamic equilibrium with small amounts of net negative and net positive ions. At the exact midpoint between the two pKa values, the trace amount of net negative and trace of net positive ions exactly balance, so that average net charge of all forms present is zero.[19] This pH is known as the isoelectric point pI, so pI = ½(pKa1 + pKa2). The individual amino acids all have slightly different pKa values, so have different isoelectric points. For amino acids with charged side-chains, the pKa of the side-chain is involved. Thus for Asp, Glu with negative side-chains, pI = ½(pKa1 + pKaR), where pKaR is the side-chain pKa. Cysteine also has potentially negative side-chain with pKaR = 8.14, so pI should be calculated as for Asp and Glu, even though the side-chain is not significantly charged at neutral pH. For His, Lys, and Arg with positive side-chains, pI = ½(pKaR + pKa2). Amino acids have zero mobility in electrophoresis at their isoelectric point, although this behaviour is more usually exploited for peptides and proteins than single amino acids. Zwitterions have minimum solubility at their isolectric point and some amino acids (in particular, with non-polar side-chains) can be isolated by precipitation from water by adjusting the pH to the required isoelectric point.

Occurrence and functions in biochemistry

A protein depicted as a long unbranched string of linked circles each representing amino acids. One circle is magnified, to show the general structure of an amino acid. This is a simplified model of the repeating structure of protein, illustrating how amino acids are joined together in these molecules.
A polypeptide is an unbranched chain of amino acids.

Standard amino acids

Amino acids are the structural units that make up proteins. They join together to form short polymer chains called peptides or longer chains called either polypeptides or proteins. These polymers are linear and unbranched, with each amino acid within the chain attached to two neighboring amino acids. 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.[20] 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-two amino acids are naturally incorporated into polypeptides and are called proteinogenic or natural amino acids.[9] Of these, 20 are encoded by the universal genetic code. The remaining 2, selenocysteine and pyrrolysine, are incorporated into proteins by unique synthetic mechanisms. Selenocysteine is incorporated when the mRNA being translated includes a SECIS element, which causes the UGA codon to encode selenocysteine instead of a stop codon.[21] Pyrrolysine is used by some methanogenic archaea in enzymes that they use to produce methane. It is coded for with the codon UAG, which is normally a stop codon in other organisms.[22] This UAG codon is followed by a PYLIS downstream sequence.[23]

The structure of selenocysteine, this differs from the lead image by having the R group (the side-chain) replaced by a carbon atom with two hydrogen and a selenium attached.
The amino acid selenocysteine

Non-standard amino acids

Aside from the 22 standard amino acids, there are many other amino acids that are called non-proteinogenic or non-standard. Those either are not found in proteins (for example carnitine, GABA), or are not produced directly and in isolation by standard cellular machinery (for example, hydroxyproline and selenomethionine).

Non-standard amino acids that are found in proteins are formed by post-translational modification, which is modification after translation during 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,[24] and the hydroxylation of proline is critical for maintaining connective tissues.[25] Another example is the formation of hypusine in the translation initiation factor EIF5A, through modification of a lysine residue.[26] 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.[27]

Comparison of the structures of alanine and beta alanine. In alanine, the side-chain is a methyl group; in beta alanine, the side-chain contains a methylene group connected to an amino group, and the alpha carbon lacks an amino group. The two amino acids, therefore, have the same formulae but different structures.
β-alanine and its α-alanine isomer

Some nonstandard amino acids are not found in proteins. Examples 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).[28] 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.[29]

In human nutrition

When taken up into the human body from the diet, the 22 standard amino acids either are used to synthesize proteins and other biomolecules or are oxidized to urea and carbon dioxide as a source of energy.[30] 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.[31] Glucogenic amino acids can also be converted into glucose, through gluconeogenesis.[32]

Pyrrolysine trait is restricted to several microbes, and only one organism has both Pyl and Sec. Of the 22 standard amino acids, 9 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.[33] In addition, 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.[34][35] 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 Nonessential
Isoleucine Alanine
Leucine Asparagine
Lysine Aspartic acid
Methionine Cysteine*
Phenylalanine Glutamic acid
Threonine Glutamine*
Tryptophan Glycine*
Valine Proline*
Histidine Selenocysteine*
Serine*
Tyrosine*
Arginine*
Ornithine*
Taurine*

(*) Essential only in certain cases.[36][37]

Non-protein functions

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:

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.[43]

Some non-standard amino acids are used as defenses against herbivores in plants.[44] For example canavanine is an analogue of arginine that is found in many legumes,[45] and in particularly large amounts in Canavalia gladiata (sword bean).[46] 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.[47] The non-protein amino acid mimosine is found in other species of legume, particularly Leucaena leucocephala.[48] This compound is an analogue of tyrosine and can poison animals that graze on these plants.

Uses in technology

Amino acids are used for a variety of applications in industry, but their main use is as additives to animal feed. This is necessary, since many of the bulk components of these feeds, such as soybeans, either have low levels or lack some of the essential amino acids: Lysine, methionine, threonine, and tryptophan are most important in the production of these feeds.[49] In this industry, amino acids are also used to chelate metal cations in order to improve the absorption of minerals from supplements, which may be required to improve the health or production of these animals.[50]

The food industry is also a major consumer of amino acids, in particular, glutamic acid, which is used as a flavor enhancer,[51] and Aspartame (aspartyl-phenylalanine-1-methyl ester) as a low-calorie artificial sweetener.[52] Similar technology to that used for animal nutrition is employed in the human nutrition industry to alleviate symptoms of mineral deficiencies, such as anemia, by improving mineral absorption and reducing negative side effects from inorganic mineral supplementation. [53]

The chelating ability of amino acids has been used in fertilizers for agriculture to facilitate the delivery of minerals to plants in order to correct mineral deficiencies, such as iron chlorosis. These fertilizers are also used to prevent deficiencies from occurring and improving the overall health of the plants.[54] The remaining production of amino acids is used in the synthesis of drugs and cosmetics.[49]

Amino acid derivative Pharmaceutical application
5-HTP (5-hydroxytryptophan) Experimental treatment for depression.[55]
L-DOPA (L-dihydroxyphenylalanine) Treatment for Parkinsonism.[56]
Eflornithine Drug that inhibits ornithine decarboxylase and is used in the treatment of sleeping sickness.[57]

Expanded genetic code

Since 2001, 40 non-natural amino acids have been added into protein by creating a unique codon (recoding) and a corresponding transfer-RNA:aminoacyl – tRNA-synthetase pair to encode it with diverse physicochemical and biological properties in order to be used as a tool to exploring protein structure and function or to create novel or enhanced proteins.[58][59]

Chemical building blocks

Amino acids are important as low-cost feedstocks. These compounds are used in chiral pool synthesis as enantiomerically-pure building blocks.[60]

Amino acids have been investigated as precursors chiral catalysts, e.g., for asymmetric hydrogenation reactions, although no commercial applications exist.[61]

Biodegradable plastics

Amino acids are under development as components of a range of biodegradable polymers. These materials have applications as environmentally friendly packaging and in medicine in drug delivery and the construction of prosthetic implants. These polymers include polypeptides, polyamides, polyesters, polysulfides, and polyurethanes with amino acids either forming part of their main chains or bonded as side-chains. These modifications alter the physical properties and reactivities of the polymers.[62] An interesting example of such materials is polyaspartate, a water-soluble biodegradable polymer that may have applications in disposable diapers and agriculture.[63] Due to its solubility and ability to chelate metal ions, polyaspartate is also being used as a biodegradeable anti-scaling agent and a corrosion inhibitor.[64][65] In addition, the aromatic amino acid tyrosine is being developed as a possible replacement for toxic phenols such as bisphenol A in the manufacture of polycarbonates.[66]

Reactions

As amino acids have both a primary amine group and a primary carboxyl group, these chemicals can undergo most of the reactions associated with these functional groups. These include nucleophilic addition, amide bond formation and imine formation for the amine group and esterification, amide bond formation and decarboxylation for the carboxylic acid group.[67] The combination of these functional groups allow amino acids to be effective polydentate ligands for metal-amino acid chelates.[68] The multiple side-chains of amino acids can also undergo chemical reactions.[69] The types of these reactions are determined by the groups on these side-chains and are, therefore, different between the various types of amino acid.

For the steps in the reaction, see the text.
The Strecker amino acid synthesis

Chemical synthesis

Several methods exist to synthesize amino acids. One of the oldest methods begins with the bromination at the α-carbon of a carboxylic acid. Nucleophilic substitution with ammonia then converts the alkyl bromide to the amino acid.[70] In alternative fashion, the Strecker amino acid synthesis involves the treatment of an aldehyde with potassium cyanide and ammonia, this produces an α-amino nitrile as an intermediate. Hydrolysis of the nitrile in acid then yields a α-amino acid.[71] Using ammonia or ammonium salts in this reaction gives unsubstituted amino acids, while substituting primary and secondary amines will yield substituted amino acids.[72] Likewise, using ketones, instead of aldehydes, gives α,α-disubstituted amino acids.[73] The classical synthesis gives racemic mixtures of α-amino acids as products, but several alternative procedures using asymmetric auxiliaries [74] or asymmetric catalysts [75][76] have been developed.[77]

At the current time, the most-adopted method is an automated synthesis on a solid support (e.g., polystyrene beads), using protecting groups (e.g., Fmoc and t-Boc) and activating groups (e.g., DCC and DIC).

Peptide bond formation

Two amino acids are shown next to each other. One loses a hydrogen and oxygen from its carboxyl group (COOH) and the other loses a hydrogen from its amino group (NH2). This reaction produces a molecule of water (H2O) and two amino acids joined by a peptide bond (-CO-NH-). The two joined amino acids are called a dipeptide.
The condensation of two amino acids to form a peptide bond

As both the amine and carboxylic acid groups of amino acids can react to form amide bonds, one amino acid molecule can react with another and become joined through an amide linkage. This polymerization of amino acids is what creates proteins. This condensation reaction yields the newly formed peptide bond and a molecule of water. In cells, this reaction does not occur directly; instead the amino acid is first activated by attachment to a transfer RNA molecule through an ester bond. This aminoacyl-tRNA is produced in an ATP-dependent reaction carried out by an aminoacyl tRNA synthetase.[78] This aminoacyl-tRNA is then a substrate for the ribosome, which catalyzes the attack of the amino group of the elongating protein chain on the ester bond.[79] As a result of this mechanism, all proteins made by ribosomes are synthesized starting at their N-terminus and moving towards their C-terminus.

However, not all peptide bonds are formed in this way. In a few cases, peptides are synthesized by specific enzymes. For example, the tripeptide glutathione is an essential part of the defenses of cells against oxidative stress. This peptide is synthesized in two steps from free amino acids.[80] In the first step gamma-glutamylcysteine synthetase condenses cysteine and glutamic acid through a peptide bond formed between the side-chain carboxyl of the glutamate (the gamma carbon of this side-chain) and the amino group of the cysteine. This dipeptide is then condensed with glycine by glutathione synthetase to form glutathione.[81]

In chemistry, peptides are synthesized by a variety of reactions. One of the most-used in solid-phase peptide synthesis uses the aromatic oxime derivatives of amino acids as activated units. These are added in sequence onto the growing peptide chain, which is attached to a solid resin support.[82] The ability to easily synthesize vast numbers of different peptides by varying the types and order of amino acids (using combinatorial chemistry) has made peptide synthesis particularly important in creating libraries of peptides for use in drug discovery through high-throughput screening.[83]

Biosynthesis

In plants, nitrogen is first assimilated into organic compounds in the form of glutamate, formed from alpha-ketoglutarate and ammonia in the mitochondrion. In order to form other amino acids, the plant uses transaminases to move the amino group to another alpha-keto carboxylic acid. For example, aspartate aminotransferase converts glutamate and oxaloacetate to alpha-ketoglutarate and aspartate.[84] Other organisms use transaminases for amino acid synthesis, too.

Nonstandard amino acids are usually formed through modifications to standard amino acids. For example, homocysteine is formed through the transsulfuration pathway or by the demethylation of methionine via the intermediate metabolite S-adenosyl methionine,[43] while hydroxyproline is made by a posttranslational modification of proline.[85]

Microorganisms and plants can synthesize many uncommon amino acids. For example, some microbes make 2-aminoisobutyric acid and lanthionine, which is a sulfide-bridged derivative of alanine. Both of these amino acids are found in peptidic lantibiotics such as alamethicin.[86] While in plants, 1-aminocyclopropane-1-carboxylic acid is a small disubstituted cyclic amino acid that is a key intermediate in the production of the plant hormone ethylene.[87]

Catabolism

Catabolism of proteinogenic amino acids. Amino acids can be classified according to the properties of their main products as either of the following:[88]
  • Glucogenic, with the products having the ability to form glucose by gluconeogenesis
  • Ketogenic, with the products not having the ability to form glucose. These products may still be used for ketogenesis or lipid synthesis.
  • Amino acids catabolized into both glucogenic and ketogenic products.

Degradation of an amino acid often involves deamination by moving its amino group to alpha-ketoglutarate, forming glutamate. This process involves transaminases, often the same as those used in amination during synthesis. In many vertebrates, the amino group is then removed through the urea cycle and is excreted in the form of urea. However, amino acid degradation can produce uric acid or ammonia instead. For example, serine dehydratase converts serine to pyruvate and ammonia.[89]

Physicochemical properties of amino acids

The 20 amino acids encoded directly by the genetic code can be divided into several groups based on their properties. Important factors are charge, hydrophilicity or hydrophobicity, size, and functional groups.[9] These properties are important for protein structure and protein–protein interactions. The water-soluble proteins tend to have their hydrophobic residues (Leu, Ile, Val, Phe, and Trp) buried in the middle of the protein, whereas hydrophilic side-chains are exposed to the aqueous solvent. The integral membrane proteins tend to have outer rings of exposed hydrophobic amino acids that anchor them into the lipid bilayer. In the case part-way between these two extremes, some peripheral membrane proteins have a patch of hydrophobic amino acids on their surface that locks onto the membrane. In similar fashion, proteins that have to bind to positively-charged molecules have surfaces rich with negatively charged amino acids like glutamate and aspartate, while proteins binding to negatively-charged molecules have surfaces rich with positively charged chains like lysine and arginine. There are different hydrophobicity scales of amino acid residues.[90]

Some amino acids have special properties such as cysteine, that can form covalent disulfide bonds to other cysteine residues, proline that forms a cycle to the polypeptide backbone, and glycine that is more flexible than other amino acids.

Many proteins undergo a range of posttranslational modifications, when additional chemical groups are attached to the amino acids in proteins. Some modifications can produce hydrophobic lipoproteins,[91] or hydrophilic glycoproteins.[92] These type of modification allow the reversible targeting of a protein to a membrane. For example, the addition and removal of the fatty acid palmitic acid to cysteine residues in some signaling proteins causes the proteins to attach and then detach from cell membranes.[93]

Table of standard amino acid abbreviations and properties

Amino Acid 3-Letter[94] 1-Letter[94] Side-chain polarity[94] Side-chain charge (pH 7.4)[94] Hydropathy index[95] Absorbance λmax(nm)[96] ε at λmax (x10−3 M−1 cm−1)[96]
Alanine Ala A nonpolar neutral 1.8
Arginine Arg R polar positive −4.5
Asparagine Asn N polar neutral −3.5
Aspartic acid Asp D polar negative −3.5
Cysteine Cys C polar neutral 2.5 250 0.3
Glutamic acid Glu E polar negative −3.5
Glutamine Gln Q polar neutral −3.5
Glycine Gly G nonpolar neutral −0.4
Histidine His H polar positive(10%)

neutral(90%)

−3.2 211 5.9
Isoleucine Ile I nonpolar neutral 4.5
Leucine Leu L nonpolar neutral 3.8
Lysine Lys K polar positive −3.9
Methionine Met M nonpolar neutral 1.9
Phenylalanine Phe F nonpolar neutral 2.8 257, 206, 188 0.2, 9.3, 60.0
Proline Pro P nonpolar neutral −1.6
Serine Ser S polar neutral −0.8
Threonine Thr T polar neutral −0.7
Tryptophan Trp W nonpolar neutral −0.9 280, 219 5.6, 47.0
Tyrosine Tyr Y polar neutral −1.3 274, 222, 193 1.4, 8.0, 48.0
Valine Val V nonpolar neutral 4.2

In addition, there are two additional amino acids that are incorporated by overriding stop codons:

21st and 22nd amino acids 3-Letter 1-Letter
Selenocysteine Sec U
Pyrrolysine Pyl O

In addition to the specific amino acid codes, placeholders are used in cases where chemical or crystallographic analysis of a peptide or protein cannot conclusively determine the identity of a residue.

Ambiguous Amino Acids 3-Letter 1-Letter
Asparagine or aspartic acid Asx B
Glutamine or glutamic acid Glx Z
Leucine or Isoleucine Xle J
Unspecified or unknown amino acid Xaa X

Unk is sometimes used instead of Xaa, but is less standard.

In addition, many non-standard amino acids have a specific code. For example, several peptide drugs, such as Bortezomib and MG132, are artificially synthesized and retain their protecting groups, which have specific codes. Bortezomib is Pyz-Phe-boroLeu, and MG132 is Z-Leu-Leu-Leu-al. To aid in the analysis of protein structure, photocrosslinking amino acid analogues are available. These include photoleucine (pLeu) and photomethionine (pMet).[97]

See also

References and notes

  1. ^ Proline is an exception to this general formula. It lacks the NH2 group because of the cyclization of the side-chain and is known as an imino acid; it falls under the category of special structured amino acids.
  2. ^ "The Structures of Life". National Institute of General Medical Sciences. http://publications.nigms.nih.gov/structlife/chapter1.html. Retrieved 2008-05-20. 
  3. ^ Vauquelin LN, Robiquet PJ (1806). "The discovery of a new plant principle in Asparagus sativus". Annales de Chimie 57: 88–93. 
  4. ^ a b Anfinsen CB, Edsall JT, Richards FM (1972). Advances in Protein Chemistry. New York: Academic Press. pp. 99, 103. ISBN 978-0-12-034226-6. 
  5. ^ Wollaston WH (1810). "On cystic oxide, a new species of urinary calculus". Philosophical Transactions of the Royal Society of London 100 (0): 223–30. doi:10.1098/rstl.1810.0015. 
  6. ^ Baumann E (1884). "Über cystin und cystein". Z Physiol Chemie 8 (4): 299–305. http://vlp.mpiwg-berlin.mpg.de/library/data/lit16533. Retrieved 28 March 2011. 
  7. ^ Braconnot HM (1820). "Sur la conversion des matières animales en nouvelles substances par le moyen de l'acide sulfurique". Ann Chim Phys Ser 2 13: 113–25. 
  8. ^ "etymonline.com entry for amino". www.etymonline.com. http://www.etymonline.com/index.php?term=amino. Retrieved 2010-07-19. 
  9. ^ a b c d e f Creighton, Thomas H. (1993). "Chapter 1". Proteins: structures and molecular properties. San Francisco: W. H. Freeman. ISBN 978-0-7167-7030-5. 
  10. ^ "Nomenclature and Symbolism for Amino Acids and Peptides". IUPAC-IUB Joint Commission on Biochemical Nomenclature. 1983. http://www.chem.qmul.ac.uk/iupac/AminoAcid/AA1n2.html. Retrieved 2008-11-17. 
  11. ^ Jodidi, S. L. (1926-03-01). "The Formol Titration of Certain Amino Acids". Journal of the American Chemical Society 48 (3): 751–753. doi:10.1021/ja01414a033. 
  12. ^ Liebecq, Claude, ed (1992). Biochemical Nomenclature and Related Documents (2nd ed.). Portland Press. pp. 39–69. ISBN 978-1-85578-005-7. 
  13. ^ Smith, Anthony D. (1997). Oxford dictionary of biochemistry and molecular biology. Oxford: Oxford University Press. pp. 535. ISBN 978-0-19-854768-6. OCLC 37616711. 
  14. ^ Pisarewicz K, Mora D, Pflueger FC, Fields GB, Marí F (May 2005). "Polypeptide chains containing D-gamma-hydroxyvaline". Journal of the American Chemical Society 127 (17): 6207–15. doi:10.1021/ja050088m. PMID 15853325. 
  15. ^ van Heijenoort J (March 2001). "Formation of the glycan chains in the synthesis of bacterial peptidoglycan". Glycobiology 11 (3): 25R–36R. doi:10.1093/glycob/11.3.25R. PMID 11320055. 
  16. ^ Wolosker H, Dumin E, Balan L, Foltyn VN (July 2008). "D-amino acids in the brain: D-serine in neurotransmission and neurodegeneration". The FEBS Journal 275 (14): 3514–26. doi:10.1111/j.1742-4658.2008.06515.x. PMID 18564180. 
  17. ^ Hatem, Salama Mohamed Ali (2006). "Gas chromatographic determination of Amino Acid Enantiomers in tobacco and bottled wines". University of Giessen. http://geb.uni-giessen.de/geb/volltexte/2006/3038/index.html. Retrieved 2008-11-17. 
  18. ^ Simmons, William J.; Gerhard Meisenberg (2006). Principles of medical biochemistry. Mosby Elsevier. p. 19. ISBN 0-323-02942-6. 
  19. ^ Fennema OR. Food Chemistry 3rd Ed. CRC Press. pp. 327–8. ISBN 0-8247-9691-8. 
  20. ^ Rodnina MV, Beringer M, Wintermeyer W (January 2007). "How ribosomes make peptide bonds". Trends in Biochemical Sciences 32 (1): 20–6. doi:10.1016/j.tibs.2006.11.007. PMID 17157507. 
  21. ^ Driscoll DM, Copeland PR (2003). "Mechanism and regulation of selenoprotein synthesis". Annual Review of Nutrition 23 (1): 17–40. doi:10.1146/annurev.nutr.23.011702.073318. PMID 12524431. 
  22. ^ Krzycki JA (December 2005). "The direct genetic encoding of pyrrolysine". Current Opinion in Microbiology 8 (6): 706–12. doi:10.1016/j.mib.2005.10.009. PMID 16256420. 
  23. ^ Théobald-Dietrich A, Giegé R, Rudinger-Thirion J (2005). "Evidence for the existence in mRNAs of a hairpin element responsible for ribosome dependent pyrrolysine insertion into proteins". Biochimie 87 (9–10): 813–7. doi:10.1016/j.biochi.2005.03.006. PMID 16164991. 
  24. ^ Vermeer C (March 1990). "Gamma-carboxyglutamate-containing proteins and the vitamin K-dependent carboxylase". The Biochemical Journal 266 (3): 625–36. PMC 1131186. PMID 2183788. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1131186. 
  25. ^ Bhattacharjee A, Bansal M (March 2005). "Collagen structure: the Madras triple helix and the current scenario". IUBMB Life 57 (3): 161–72. doi:10.1080/15216540500090710. PMID 16036578. 
  26. ^ Park MH (February 2006). "The post-translational synthesis of a polyamine-derived amino acid, hypusine, in the eukaryotic translation initiation factor 5A (eIF5A)". Journal of Biochemistry 139 (2): 161–9. doi:10.1093/jb/mvj034. PMC 2494880. PMID 16452303. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2494880. 
  27. ^ Blenis J, Resh MD (December 1993). "Subcellular localization specified by protein acylation and phosphorylation". Current Opinion in Cell Biology 5 (6): 984–9. doi:10.1016/0955-0674(93)90081-Z. PMID 8129952. 
  28. ^ Curis E, Nicolis I, Moinard C, et al. (November 2005). "Almost all about citrulline in mammals". Amino Acids 29 (3): 177–205. doi:10.1007/s00726-005-0235-4. PMID 16082501. 
  29. ^ Coxon KM, Chakauya E, Ottenhof HH, et al. (August 2005). "Pantothenate biosynthesis in higher plants". Biochemical Society Transactions 33 (Pt 4): 743–6. doi:10.1042/BST0330743. PMID 16042590. 
  30. ^ Sakami W, Harrington H (1963). "Amino acid metabolism". Annual Review of Biochemistry 32 (1): 355–98. doi:10.1146/annurev.bi.32.070163.002035. PMID 14144484. 
  31. ^ Brosnan JT (April 2000). "Glutamate, at the interface between amino acid and carbohydrate metabolism". The Journal of Nutrition 130 (4S Suppl): 988S–90S. PMID 10736367. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=10736367. 
  32. ^ Young VR, Ajami AM (September 2001). "Glutamine: the emperor or his clothes?". The Journal of Nutrition 131 (9 Suppl): 2449S–59S; discussion 2486S–7S. PMID 11533293. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=11533293. 
  33. ^ Young VR (August 1994). "Adult amino acid requirements: the case for a major revision in current recommendations". The Journal of Nutrition 124 (8 Suppl): 1517S–1523S. PMID 8064412. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=8064412. 
  34. ^ Imura K, Okada A (January 1998). "Amino acid metabolism in pediatric patients". Nutrition 14 (1): 143–8. doi:10.1016/S0899-9007(97)00230-X. PMID 9437700. 
  35. ^ Lourenço R, Camilo ME (2002). "Taurine: a conditionally essential amino acid in humans? An overview in health and disease". Nutrición Hospitalaria 17 (6): 262–70. PMID 12514918. 
  36. ^ Fürst P, Stehle P (June 2004). "What are the essential elements needed for the determination of amino acid requirements in humans?". The Journal of Nutrition 134 (6 Suppl): 1558S–1565S. PMID 15173430. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=15173430. 
  37. ^ Reeds PJ (July 2000). "Dispensable and indispensable amino acids for humans". The Journal of Nutrition 130 (7): 1835S–40S. PMID 10867060. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=10867060. 
  38. ^ Savelieva KV, Zhao S, Pogorelov VM, et al. (2008). Bartolomucci, Alessandro. ed. "Genetic disruption of both tryptophan hydroxylase genes dramatically reduces serotonin and affects behavior in models sensitive to antidepressants". PloS ONE 3 (10): e3301. Bibcode 2008PLoSO...3.3301S. doi:10.1371/journal.pone.0003301. PMC 2565062. PMID 18923670. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2565062. 
  39. ^ Shemin D, Rittenberg D (1 December 1946). "The biological utilization of glycine for the synthesis of the protoporphyrin of hemoglobin". Journal of Biological Chemistry 166 (2): 621–5. PMID 20276176. http://www.jbc.org/cgi/reprint/166/2/621. 
  40. ^ Tejero J, Biswas A, Wang ZQ, et al. (November 2008). "Stabilization and characterization of a heme-oxy reaction intermediate in inducible nitric-oxide synthase". The Journal of Biological Chemistry 283 (48): 33498–507. doi:10.1074/jbc.M806122200. PMC 2586280. PMID 18815130. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2586280. 
  41. ^ Rodríguez-Caso C, Montañez R, Cascante M, Sánchez-Jiménez F, Medina MA (August 2006). "Mathematical modeling of polyamine metabolism in mammals". The Journal of Biological Chemistry 281 (31): 21799–812. doi:10.1074/jbc.M602756200. PMID 16709566. 
  42. ^ Stryer, Lubert; Berg, Jeremy Mark; Tymoczko, John L. (2002). Biochemistry. San Francisco: W.H. Freeman. pp. 693–8. ISBN 0-7167-4684-0. 
  43. ^ a b Brosnan JT, Brosnan ME (June 2006). "The sulfur-containing amino acids: an overview". The Journal of Nutrition 136 (6 Suppl): 1636S–1640S. PMID 16702333. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=16702333. 
  44. ^ Hylin, John W. (1969). "Toxic peptides and amino acids in foods and feeds". Journal of Agricultural and Food Chemistry 17 (3): 492–6. doi:10.1021/jf60163a003. 
  45. ^ Turner, B. L.; Harborne, J. B. (1967). "Distribution of canavanine in the plant kingdom". Phytochemistry 6 (6): 863–66. doi:10.1016/S0031-9422(00)86033-1. 
  46. ^ Ekanayake S, Skog K, Asp NG (May 2007). "Canavanine content in sword beans (Canavalia gladiata): analysis and effect of processing". Food and Chemical Toxicology 45 (5): 797–803. doi:10.1016/j.fct.2006.10.030. PMID 17187914. 
  47. ^ Rosenthal GA (2001). "L-Canavanine: a higher plant insecticidal allelochemical". Amino Acids 21 (3): 319–30. doi:10.1007/s007260170017. PMID 11764412. 
  48. ^ Hammond AC (May 1995). "Leucaena toxicosis and its control in ruminants". Journal of Animal Science 73 (5): 1487–92. PMID 7665380. http://jas.fass.org/cgi/pmidlookup?view=long&pmid=7665380. 
  49. ^ a b Leuchtenberger W, Huthmacher K, Drauz K (November 2005). "Biotechnological production of amino acids and derivatives: current status and prospects". Applied Microbiology and Biotechnology 69 (1): 1–8. doi:10.1007/s00253-005-0155-y. PMID 16195792. 
  50. ^ Ashmead, H. DeWayne (1993). The Role of Amino Acid Chelates in Animal Nutrition. Westwood: Noyes Publications. 
  51. ^ Garattini S (April 2000). "Glutamic acid, twenty years later". The Journal of Nutrition 130 (4S Suppl): 901S–9S. PMID 10736350. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=10736350. 
  52. ^ Stegink LD (July 1987). "The aspartame story: a model for the clinical testing of a food additive". The American Journal of Clinical Nutrition 46 (1 Suppl): 204–15. PMID 3300262. http://www.ajcn.org/cgi/pmidlookup?view=long&pmid=3300262. 
  53. ^ Albion Laboratories, Inc.. "Albion Ferrochel Website". http://www.albionferrochel.com. Retrieved Cited: July 12, 2011. 
  54. ^ Ashmead, H. DeWayne (1986). Foliar Feeding of Plants with Amino Acid Chelates. Park Ridge: Noyes Publications. 
  55. ^ Turner EH, Loftis JM, Blackwell AD (March 2006). "Serotonin a la carte: supplementation with the serotonin precursor 5-hydroxytryptophan". Pharmacology & Therapeutics 109 (3): 325–38. doi:10.1016/j.pharmthera.2005.06.004. PMID 16023217. 
  56. ^ Kostrzewa RM, Nowak P, Kostrzewa JP, Kostrzewa RA, Brus R (March 2005). "Peculiarities of L: -DOPA treatment of Parkinson's disease". Amino Acids 28 (2): 157–64. doi:10.1007/s00726-005-0162-4. PMID 15750845. 
  57. ^ Heby O, Persson L, Rentala M (August 2007). "Targeting the polyamine biosynthetic enzymes: a promising approach to therapy of African sleeping sickness, Chagas' disease, and leishmaniasis". Amino Acids 33 (2): 359–66. doi:10.1007/s00726-007-0537-9. PMID 17610127. 
  58. ^ Xie J, Schultz PG (December 2005). "Adding amino acids to the genetic repertoire". Curr Opin Chem Biol 9 (6): 548–54. doi:10.1016/j.cbpa.2005.10.011. PMID 16260173. 
  59. ^ Wang Q, Parrish AR, Wang L (March 2009). "Expanding the genetic code for biological studies". Chem. Biol. 16 (3): 323–36. doi:10.1016/j.chembiol.2009.03.001. PMC 2696486. PMID 19318213. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2696486. 
  60. ^ Hanessian, S. (1993). "Reflections on the total synthesis of natural products: Art, craft, logic, and the chiron approach". Pure and Applied Chemistry 65 (6): 1189–204. doi:10.1351/pac199365061189. 
  61. ^ Blaser, Hans Ulrich (1992). "The chiral pool as a source of enantioselective catalysts and auxiliaries". Chemical Reviews 92 (5): 935–52. doi:10.1021/cr00013a009. 
  62. ^ Sanda, Fumio; Endo, Takeshi (1999). "Feature Article Syntheses and functions of polymers based on amino acids". Macromolecular Chemistry and Physics 200 (12): 2651–61. doi:10.1002/(SICI)1521-3935(19991201)200:12<2651::AID-MACP2651>3.0.CO;2-P. 
  63. ^ Gross, R. A.; Kalra, B. (2002). "Biodegradable Polymers for the Environment". Science 297 (5582): 803–807. Bibcode 2002Sci...297..803G. doi:10.1126/science.297.5582.803. PMID 12161646. http://www.sciencemag.org/cgi/content/abstract/297/5582/803. 
  64. ^ Low, K. C.; Wheeler, A. P.; Koskan, L. P. (1996). Commercial poly(aspartic acid) and Its Uses. Advances in Chemistry Series. 248. Washington, D.C.: American Chemical Society. 
  65. ^ Thombre, S.M.; Sarwade, B.D. (2005). "Synthesis and Biodegradability of Polyaspartic Acid: A Critical Review". Journal of Macromolecular Science, Part A 42 (9): 1299–1315. doi:10.1080/10601320500189604. http://www.informaworld.com/index/718581646.pdf. 
  66. ^ Bourke, S. L.; Kohn, J. (2003). "Polymers derived from the amino acid l-tyrosine: polycarbonates, polyarylates and copolymers with poly(ethylene glycol)". Advanced Drug Delivery Reviews 55 (4): 447–466. doi:10.1016/S0169-409X(03)00038-3. PMID 12706045. http://linkinghub.elsevier.com/retrieve/pii/S0169409X03000383. 
  67. ^ Elmore, Donald Trevor; Barrett, G. C. (1998). Amino acids and peptides. Cambridge, UK: Cambridge University Press. pp. 48–60. ISBN 0-521-46827-2. 
  68. ^ Konar, Sanjit; et al. (2010). "Structural determination and characterization of copper and zinc bis-glycinates with X-ray crystallography and mass spectrometry". Journal of Coordination Chemistry 63 (19). 
  69. ^ Gutteridge A, Thornton JM (November 2005). "Understanding nature's catalytic toolkit". Trends in Biochemical Sciences 30 (11): 622–9. doi:10.1016/j.tibs.2005.09.006. PMID 16214343. 
  70. ^ McMurry, John (1996). Organic chemistry. Pacific Grove, CA, USA: Brooks/Cole. p. 1064. ISBN 0-534-23832-7. 
  71. ^ Strecker, Adolph (1850). "Ueber die künstliche Bildung der Milchsäure und einen neuen, dem Glycocoll homologen Körper". Justus Liebigs Annalen der Chemie 75 (1): 27–45. doi:10.1002/jlac.18500750103. 
  72. ^ Strecker, Adolph (1854). "Ueber einen neuen aus Aldehyd - Ammoniak und Blausäure entstehenden Körper". Justus Liebigs Annalen der Chemie 91 (3): 349–51. doi:10.1002/jlac.18540910309. 
  73. ^ Masumoto S, Usuda H, Suzuki M, Kanai M, Shibasaki M (May 2003). "Catalytic enantioselective Strecker reaction of ketoimines". Journal of the American Chemical Society 125 (19): 5634–5. doi:10.1021/ja034980. PMID 12733893. 
  74. ^ Davis, F. A.; Reddy, Rajarathnam E.; Portonovo, Padma S. (1994). "Asymmetric strecker synthesis using enantiopure sulfinimines: A convenient synthesis of α-amino acids". Tetrahedron Letters 35 (50): 9351. doi:10.1016/S0040-4039(00)78540-6. 
  75. ^ Ishitani, Haruro; Komiyama, Susumu; Hasegawa, Yoshiki; Kobayashi, Shū (2000). "Catalytic Asymmetric Strecker Synthesis. Preparation of Enantiomerically Pure α-Amino Acid Derivatives from Aldimines and Tributyltin Cyanide or Achiral Aldehydes, Amines, and Hydrogen Cyanide Using a Chiral Zirconium Catalyst". Journal of the American Chemical Society 122 (5): 762–6. doi:10.1021/ja9935207. 
  76. ^ Huang, Jinkun; Corey, E. J. (2004). "A New Chiral Catalyst for the Enantioselective Strecker Synthesis of α-Amino Acids". Orgic Letters 62 (6): 5027–9. doi:10.1021/ol047698w. PMID 15606127. 
  77. ^ Duthaler, Rudolf O. (1994). "Recent developments in the stereoselective synthesis of α-aminoacids". Tetrahedron 50 (6): 1539–1650. doi:10.1016/S0040-4020(01)80840-1. 
  78. ^ Ibba M, Söll D (May 2001). "The renaissance of aminoacyl-tRNA synthesis". EMBO Reports 2 (5): 382–7. doi:10.1093/embo-reports/kve095 (inactive 2010-02-18). PMC 1083889. PMID 11375928. http://www.nature.com/embor/journal/v2/n5/full/embor420.html. 
  79. ^ Lengyel P, Söll D (June 1969). "Mechanism of protein biosynthesis". Bacteriological Reviews 33 (2): 264–301. PMC 378322. PMID 4896351. http://mmbr.asm.org/cgi/pmidlookup?view=long&pmid=4896351. 
  80. ^ Wu G, Fang YZ, Yang S, Lupton JR, Turner ND (March 2004). "Glutathione metabolism and its implications for health". The Journal of Nutrition 134 (3): 489–92. PMID 14988435. http://jn.nutrition.org/cgi/pmidlookup?view=long&pmid=14988435. 
  81. ^ Meister A (November 1988). "Glutathione metabolism and its selective modification". The Journal of Biological Chemistry 263 (33): 17205–8. PMID 3053703. http://www.jbc.org/cgi/pmidlookup?view=long&pmid=3053703. 
  82. ^ Carpino, Louis A. (1992). "1-Hydroxy-7-azabenzotriazole. An efficient peptide coupling additive". Journal of the American Chemical Society 115 (10): 4397–8. doi:10.1021/ja00063a082. 
  83. ^ Marasco D, Perretta G, Sabatella M, Ruvo M (October 2008). "Past and future perspectives of synthetic peptide libraries". Current Protein & Peptide Science 9 (5): 447–67. doi:10.2174/138920308785915209. PMID 18855697. 
  84. ^ Jones, Russell Celyn; Buchanan, Bob B.; Gruissem, Wilhelm (2000). Biochemistry & molecular biology of plants. Rockville, Md: American Society of Plant Physiologists. pp. 371–2. ISBN 0-943088-39-9. 
  85. ^ Kivirikko KI, Pihlajaniemi T (1998). "Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4-hydroxylases". Advances in Enzymology and Related Areas of Molecular Biology 72: 325–98. PMID 9559057. 
  86. ^ Whitmore L, Wallace BA (May 2004). "Analysis of peptaibol sequence composition: implications for in vivo synthesis and channel formation". European Biophysics Journal 33 (3): 233–7. doi:10.1007/s00249-003-0348-1. PMID 14534753. 
  87. ^ Alexander L, Grierson D (October 2002). "Ethylene biosynthesis and action in tomato: a model for climacteric fruit ripening". Journal of Experimental Botany 53 (377): 2039–55. doi:10.1093/jxb/erf072. PMID 12324528. 
  88. ^ Chapter 20 (Amino Acid Degradation and Synthesis) in: Denise R., PhD. Ferrier. Lippincott's Illustrated Reviews: Biochemistry (Lippincott's Illustrated Reviews). Hagerstwon, MD: Lippincott Williams & Wilkins. ISBN 0-7817-2265-9. 
  89. ^ Stryer, Lubert; Berg, Jeremy Mark; Tymoczko, John L. (2002). Biochemistry. San Francisco: W.H. Freeman. pp. 639–49. ISBN 0-7167-4684-0. 
  90. ^ Urry, Dan W. (2004). "The change in Gibbs free energy for hydrophobic association: Derivation and evaluation by means of inverse temperature transitions". Chemical Physics Letters 399 (1–3): 177–83. doi:10.1016/S0009-2614(04)01565-9. 
  91. ^ Magee T, Seabra MC (April 2005). "Fatty acylation and prenylation of proteins: what's hot in fat". Current Opinion in Cell Biology 17 (2): 190–6. doi:10.1016/j.ceb.2005.02.003. PMID 15780596. 
  92. ^ Pilobello KT, Mahal LK (June 2007). "Deciphering the glycocode: the complexity and analytical challenge of glycomics". Current Opinion in Chemical Biology 11 (3): 300–5. doi:10.1016/j.cbpa.2007.05.002. PMID 17500024. 
  93. ^ Smotrys JE, Linder ME (2004). "Palmitoylation of intracellular signaling proteins: regulation and function". Annual Review of Biochemistry 73 (1): 559–87. doi:10.1146/annurev.biochem.73.011303.073954. PMID 15189153. 
  94. ^ a b c d Hausman, Robert E.; Cooper, Geoffrey M. (2004). The cell: a molecular approach. Washington, D.C: ASM Press. p. 51. ISBN 0-87893-214-3. 
  95. ^ Kyte J, Doolittle RF (May 1982). "A simple method for displaying the hydropathic character of a protein". Journal of Molecular Biology 157 (1): 105–32. doi:10.1016/0022-2836(82)90515-0. PMID 7108955. 
  96. ^ a b Freifelder, D. (1983). Physical Biochemistry (2nd ed.). W. H. Freeman and Company. ISBN 0-7167-1315-2. [page needed]
  97. ^ Suchanek M, Radzikowska A, Thiele C (April 2005). "Photo-leucine and photo-methionine allow identification of protein-protein interactions in living cells". Nature Methods 2 (4): 261–7. doi:10.1038/nmeth752. PMID 15782218. 

Further reading

  • Doolittle, R.F. (1989) Redundancies in protein sequences. In Predictions of Protein Structure and the Principles of Protein Conformation (Fasman, G.D. ed) Plenum Press, New York, pp. 599–623
  • David L. Nelson and Michael M. Cox, Lehninger Principles of Biochemistry, 3rd edition, 2000, Worth Publishers, ISBN 1-57259-153-6
  • Meierhenrich, U.J.: Amino acids and the asymmetry of life, Springer-Verlag, Berlin, New York, 2008. ISBN 978-3-540-76885-2
  • Morelli, Robert J. "Studies of amino acid absorption from the small intestine." San Francisco: Morelli, 1952.

External links


Wikimedia Foundation. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • amino acid — n. 1. any of a large group of organic acids containing a carboxyl group, COOH, and an amino group, NH2 2. any of the 25 amino acids that link together into polypeptide chains to form proteins that are necessary for all life: in general, they are… …   English World dictionary

  • amino acid — ► NOUN ▪ any of a class of about twenty organic compounds which form the basic constituents of proteins and contain both acid and amine groups …   English terms dictionary

  • amino acid — amino acid. См. аминокислота. (Источник: «Англо русский толковый словарь генетических терминов». Арефьев В.А., Лисовенко Л.А., Москва: Изд во ВНИРО, 1995 г.) …   Молекулярная биология и генетика. Толковый словарь.

  • amino acid —  Amino Acid  Аминокислота   Органическое соединение, в молекуле которого одновременно содержатся карбоксильная и аминная группы; в 20 природных аминокислотах обе группы связаны с одним и тем же атомом C.   Аминокислоты объединяются амидными… …   Толковый англо-русский словарь по нанотехнологии. - М.

  • amino acid — Biochem. any of a class of organic compounds that contains at least one amino group, NH2, and one carboxyl group, COOH: the alpha amino acids, RCH(NH2)COOH, are the building blocks from which proteins are constructed. Cf. essential amino acid.… …   Universalium

  • Amino acid — One of the 20 building blocks of protein. The sequence of amino acids in a protein and, hence, the function of that protein are determined by the genetic code in the DNA. Amino acids are molecules that (in technical terms) contain a basic amino… …   Medical dictionary

  • amino acid — n. Biochem. any of a group of organic compounds containing both the carboxyl (COOH) and amino (NH2) group, occurring naturally in plant and animal tissues and forming the basic constituents of proteins. Etymology: AMINE + ACID * * * amino acid… …   Useful english dictionary

  • amino acid — UK [əˌmiːnəʊ ˈæsɪd] / US [əˌmɪnoʊ ˈæsɪd] noun [countable] Word forms amino acid : singular amino acid plural amino acids biology one of the substances in the body that combine to make proteins …   English dictionary

  • amino acid — A compound containing both amino ( NH2) and carboxyl ( COOH) groups. In particular, any of 20 basic building blocks of proteins having the formula NH2 CR COOH, where R is different for each specific amino acid. See: annex 3 …   Glossary of Biotechnology

  • amino acid — noun Amino acid is used before these nouns: ↑sequence …   Collocations dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”