Sunteți pe pagina 1din 18

Lysine

Lysine (symbol Lys or K)[1] is an α-amino acid that is used in the


Lysine
biosynthesis of proteins. It contains an α-amino group (which is
in the protonated −NH3+ form under biological conditions), an α-
carboxylic acid group (which is in the deprotonated −COO− form
under biological conditions), and a side chain lysyl ((CH2)4NH2),
classifying it as a basic, charged (at physiological pH), aliphatic
amino acid. It is encoded by the codons, AAA and AAG. Like L-lysine
almost all other amino acids, the α-carbon is chiral and lysine Names
may refer to either enantiomer or a racemic mixture of both. For
IUPAC name
the purpose of this article, lysine will refer to the biologically
(2S)-2,6-Diaminohexanoic acid
active enantiomer L-lysine, where the α-carbon is in the S
(L-lysine) (2R)-2,6-
configuration.
Diaminohexanoic acid (D-lysine)
The human body cannot synthesize lysine, so it is essential in Other names
humans and must be obtained from the diet. In organisms that Lysine, D-lysine, L-lysine, LYS,
synthesise lysine, it has two main biosynthetic pathways, the h-Lys-OH
diaminopimelate and α-aminoadipate pathways, which employ Identifiers
different enzymes and substrates and are found in different
CAS Number 70-54-2 (http://ww
organisms. Lysine catabolism occurs through one of several
w.commonchemist
pathways, the most common of which is the saccharopine
ry.org/ChemicalDe
pathway.
tail.aspx?ref=70-5
Lysine plays several roles in humans, most importantly 4-2) DL
proteinogenesis, but also in the crosslinking of collagen 56-87-1 (http://ww
polypeptides, uptake of essential mineral nutrients, and in the w.commonchemist
production of carnitine, which is key in fatty acid metabolism. ry.org/ChemicalDe
Lysine is also often involved in histone modifications, and thus, tail.aspx?ref=56-8
impacts the epigenome. The ε-amino group often participates in 7-1) L
hydrogen bonding and as a general base in catalysis. The ε- 923-27-3 (http://w
ammonium group (NH3+) is attached to the fourth carbon from ww.commonchemi
the α-carbon, which is attached to the carboxyl (C=OOH) stry.org/Chemical
group.[2] Detail.aspx?ref=9
23-27-3) D
Due to its importance in several biological processes, a lack of
lysine can lead to several disease states including defective 3D model Interactive image
(JSmol)
connective tissues, impaired fatty acid metabolism, anaemia, and (https://chemapps.
systemic protein-energy deficiency. In contrast, an stolaf.edu/jmol/jm
overabundance of lysine, caused by ineffective catabolism, can ol.php?model=NC
cause severe neurological issues. CCCC%28N%29
C%28%3DO%29
Lysine was first isolated by the German biological chemist O)
Ferdinand Heinrich Edmund Drechsel in 1889 from the protein
Zwitterion:
casein in milk.[3] He named it "lysin".[4] In 1902, the German
Interactive image
chemists Emil Fischer and Fritz Weigert determined lysine's (https://chemapps.
chemical structure by synthesizing it.[5] stolaf.edu/jmol/jm
ol.php?model=NC
CCCC%28%5BN
H3%2B%5D%29
Contents C%28%3DO%2
Biosynthesis 9%5BO-%5D)

Catabolism Protonated
zwitterion:
Nutritional value
Interactive image
Dietary sources
(https://chemapps.
Biological roles stolaf.edu/jmol/jm
Disputed roles ol.php?model=%5
Roles in disease BNH3%2B%5DC
Use of lysine in animal feed CCCC%28%5BN
H3%2B%5D%29
In popular culture
C%28%3DO%2
References 9%5BO-%5D)
ChEBI CHEBI:25094 (htt
Biosynthesis ps://www.ebi.ac.u
k/chebi/searchId.d
Two different pathways o?chebiId=25094)
have been identified in
nature for the synthesis of
ChEMBL ChEMBL28328 (ht
lysine. The
tps://www.ebi.ac.u
diaminopimelate (DAP)
k/chembldb/index.
pathway belongs to the
php/compound/ins
aspartate derived
pect/ChEMBL283
biosynthetic family,
28)
which is also involved in
the synthesis of threonine, ChemSpider 843 (http://www.ch
methionine and emspider.com/Ch
isoleucine. [6][7] Whereas emical-Structure.8
the α-aminoadipate 43.html)
(AAA) pathway is part of 5747 (http://www.c
the glutamate hemspider.com/C
Lysine biosynthesis pathways.
biosynthetic family.[8][9] hemical-Structure.
Two pathways are responsible for the
de novo biosynthesis of L-lysine, 5747.html) L
namely the (A) diaminopimelate
The DAP pathway is
found in both prokaryotes ECHA InfoCard 100.000.673 (http
pathway and (B) α‑aminoadipate
pathway. and plants and begins s://echa.europa.e
with the u/substance-infor
dihydrodipicolinate mation/-/substanc
synthase (DHDPS) (E.C 4.3.3.7) catalysed condensation reaction einfo/100.000.67
between the aspartate derived, L-aspartate semialdehyde, and 3)
pyruvate to form (4S)-4-hydroxy-2,3,4,5-tetrahydro-(2S)- IUPHAR/BPS 724 (http://www.gu
dipicolinic acid (HTPA).[10][11][12][13][14] The product is then idetopharmacolog
reduced by dihydrodipicolinate reductase (DHDPR) (E.C y.org/GRAC/Ligan
1.3.1.26), with NAD(P)H as a proton donor, to yield 2,3,4,5- dDisplayForward?
tetrahydrodipicolinate (THDP).[15] From this point on, there are tab=summary&lig
four pathway variations found in different species, namely the andId=724)
acetylase, aminotransferase, dehydrogenase, and succinylase KEGG C16440 (https://w
pathways.[6][16] Both the acetylase and succinylase variant
ww.kegg.jp/entry/
pathways use four enzyme catalysed steps, the aminotransferase
C16440)
pathway uses two enzymes, and the dehydrogenase pathway uses
a single enzyme.[17] These four variant pathways converge at the PubChem CID 866 (https://pubch
formation of the penultimate product, meso‑diaminopimelate, em.ncbi.nlm.nih.g
which is subsequently enzymatically decarboxylated in an ov/compound/866)
irreversible reaction catalysed by diaminopimelate decarboxylase UNII K3Z4F929H6 (http
(DAPDC) (E.C 4.1.1.20) to produce L-lysine.[18][19] The DAP s://fdasis.nlm.nih.g
pathway is regulated at multiple levels, including upstream at the ov/srs/srsdirect.js
enzymes involved in aspartate processing as well as at the initial p?regno=K3Z4F9
DHDPS catalysed condensation step.[19][20] Lysine imparts a 29H6)
strong negative feedback loop on these enzymes and,
InChI
subsequently, regulates the entire pathway.[20]
InChI=1S/C6H14N2O2/c7-4-2-1-3-5(8)6(9)1
0/h5H,1-4,7-8H2,(H,9,10)
The AAA pathway involves the condensation of α-ketoglutarate Key: KDXKERNSBIXSRK-UHFFFAOYSA-N
and acetyl-CoA via the intermediate AAA for the synthesis of L-
lysine. This pathway has been shown to be present in several InChI=1/C6H14N2O2/c7-4-2-1-3-5(8)6(9)10/h
5H,1-4,7-8H2,(H,9,10)
yeast species, as well as protists and higher Key: KDXKERNSBIXSRK-UHFFFAOYAY
fungi.[9][21][22][23][24][25][26] It has also been reported that an
SMILES
alternative variant of the AAA route has been found in Thermus
NCCCCC(N)C(=O)O
thermophilus and Pyrococcus horikoshii, which could indicate
Zwitterion: NCCCCC([NH3+])C(=O)[O-]
that this pathway is more widely spread in prokaryotes than
originally proposed.[27][28][29] The first and rate-limiting step in Protonated zwitterion: [NH3+]CCCCC([NH3+])
C(=O)[O-]
the AAA pathway is the condensation reaction between acetyl-
CoA and α‑ketoglutarate catalysed by homocitrate-synthase Properties
(HCS) (E.C 2.3.3.14) to give the intermediate homocitryl‑CoA, Chemical C6H14N2O2
formula
which is hydrolysed by the same enzyme to produce
homocitrate.[30] Homocitrate is enzymatically dehydrated by Molar mass 146.190 g·mol−1
homoaconitase (HAc) (E.C 4.2.1.36) to yield cis- Solubility in 1.5 kg/L
homoaconitate.[31] HAc then catalyses a second reaction in which water
cis-homoaconitate undergoes rehydration to produce Pharmacology
[9]
homoisocitrate. The resulting product undergoes an oxidative ATC code B05XB03 (WHO
decarboxylation by homoisocitrate dehydrogenase (HIDH) (E.C (https://www.whoc
1.1.1.87) to yield α‑ketoadipate.[9] AAA is then formed via a c.no/atc_ddd_inde
pyridoxal 5′-phosphate (PLP)-dependent aminotransferase (PLP- x/?code=B05XB0
AT) (E.C 2.6.1.39), using glutamate as the amino donor.[30] From 3))
this point on, the AAA pathway differs depending on the
Supplementary data page
kingdom. In fungi, AAA is reduced to α‑aminoadipate-
Structure and Refractive index
semialdehyde via AAA reductase (E.C 1.2.1.95) in a unique
properties (n),
process involving both adenylation and reduction that is activated
by a phosphopantetheinyl transferase (E.C 2.7.8.7).[9] Once the Dielectric constant
semialdehyde is formed, saccharopine reductase (E.C 1.5.1.10) (εr), etc.
Thermodynamic
catalyses a condensation reaction with glutamate and NAD(P)H, data Phase behaviour
as a proton donor, and the imine is reduced to produce the solid–liquid–gas
penultimate product, saccharopine.[29] The final step of the Spectral data UV, IR, NMR, MS
pathway in fungi involves the saccharopine dehydrogenase Except where otherwise noted, data
(SDH) (E.C 1.5.1.8) catalysed oxidative deamination of are given for materials in their
saccharopine, resulting in L-lysine.[9] In a variant AAA pathway standard state (at 25 °C [77 °F],
found in some prokaryotes, AAA is first converted to N‑acetyl-α- 100 kPa).
aminoadipate, which is phosphorylated and then reductively verify (what is ?)
dephosphorylated to the ε-aldehyde.[29][30] The aldehyde is then Infobox references
transaminated to N‑acetyl-lysine, which is deacetylated to give L-
lysine.[29][30] However, the enzymes involved in this variant
pathway need further validation.

Catabolism
Like all amino acids, catabolism of lysine is initiated from the uptake of
dietary lysine or from the breakdown of intracellular protein.
Catabolism is also used as a means to control the intracellular
concentration of free lysine and maintain a steady-state to prevent the
toxic effects of excessive free lysine.[32] There are several pathways
involved in lysine catabolism but the most commonly used is the
saccharopine pathway, which primarily takes place in the liver (and
equivalent organs) in animals, specifically within the
mitochondria. [33][32][34][35] This is the reverse of the previously
described AAA pathway.[33][36] In animals and plants, the first two
steps of the saccharopine pathway are catalysed by the bifunctional
enzyme, α-aminoadipic semialdehyde synthase (AASS), which possess
both lysine-ketoglutarate reductase (LKR) (E.C 1.5.1.8) and SDH
activities, whereas in other organisms, such as bacteria and fungi, both
of these enzymes are encoded by separate genes.[37][38] The first step
involves the LKR catalysed reduction of L-lysine in the presence of α-
ketoglutarate to produce saccharopine, with NAD(P)H acting as a
proton donor.[39] Saccharopine then undergoes a dehydration reaction,
catalysed by SDH in the presence of NAD+, to produce AAS and
glutamate.[40] AAS dehydrogenase (AASD) (E.C 1.2.1.31) then further
dehydrates the molecule into AAA.[39] Subsequently, PLP-AT catalyses
the reverse reaction to that of the AAA biosynthesis pathway, resulting
in AAA being converted to α-ketoadipate. The product, α‑ketoadipate,
Saccharopine lysine
is decarboxylated in the presence of NAD+ and coenzyme A to yield
catabolism pathway. The
glutaryl-CoA, however the enzyme involved in this is yet to be fully saccharopine pathway is the
elucidated.[41][42] Some evidence suggests that the 2-oxoadipate most prominent pathway for
dehydrogenase complex (OADHc), which is structurally homologous to the catabolism of lysine.
the E1 subunit of the oxoglutarate dehydrogenase complex (OGDHc)
(E.C 1.2.4.2), is responsible for the decarboxylation reaction.[41][43]
Finally, glutaryl-CoA is oxidatively decarboxylated to crotony-CoA by glutaryl-CoA dehydrogenase
(E.C 1.3.8.6), which goes on to be further processed through multiple enzymatic steps to yield acetyl-
CoA; an essential carbon metabolite involved in the tricarboxylic acid cycle (TCA).[39][44][45][46]
Nutritional value
Lysine is one of the nine essential amino acids in humans.[47] The human nutritional requirements varies
from ~60 mg·kg−1·d−1 in infancy to ~30 mg·kg−1·d−1 in adults.[33] This requirement is commonly met in
a western society with the intake of lysine from meat and vegetable sources well in excess of the
recommended requirement.[33] In vegetarian diets, the intake of lysine is less due to the limiting quantity
of lysine in cereal crops compared to meat sources.[33] Given the limiting concentration of lysine in
cereal crops, it has long been speculated that the content of lysine can be increased through genetic
modification practices.[48][49] Often these practices have involved the intentional dysregulation of the
DAP pathway by means of introducing lysine feedback-insensitive orthologues of the DHDPS
enzyme.[48][49] These methods have met limited success likely due to the toxic side effects of increased
free lysine and indirect effects on the TCA cycle.[50] Plants accumulate lysine and other amino acids in
the form of seed storage proteins, found within the seeds of the plant, and this represents the edible
component of cereal crops.[51] This highlights the need to not only increase free lysine, but also direct
lysine towards the synthesis of stable seed storage proteins, and subsequently, increase the nutritional
value of the consumable component of crops.[52][53] Whilst genetic modification practices have met
limited success, more traditional selective breeding techniques have allowed for the isolation of "Quality
Protein Maize", which has significantly increased levels of lysine and tryptophan, also an essential amino
acid. This increase in lysine content is attributed to an opaque-2 mutation that reduced the transcription
of lysine lacking zein related seed storage proteins and, as a result, increased the abundance of other
proteins that are rich in lysine.[53][54] Commonly, to overcome the limiting abundance of lysine in
livestock feed, industrially produced lysine is added.[55][56] The industrial process includes the
fermentative culturing of Corynebacterium glutamicum and the subsequent purification of lysine.[55]

Dietary sources
Good sources of lysine are high-protein foods such as eggs, meat (specifically red meat, lamb, pork, and
poultry), soy, beans and peas, cheese (particularly Parmesan), and certain fish (such as cod and
sardines).[57] Lysine is the limiting amino acid (the essential amino acid found in the smallest quantity in
the particular foodstuff) in most cereal grains, but is plentiful in most pulses (legumes).[58] A vegetarian
or low animal protein diet can be adequate for protein, including lysine, if it includes both cereal grains
and legumes, but there is no need for the two food groups to be consumed in the same meals.

A food is considered to have sufficient lysine if it has at least 51 mg of lysine per gram of protein (so that
the protein is 5.1% lysine).[59] L-lysine HCl is used as a dietary supplement, providing 80.03% L-
lysine.[60] As such, 1 g of L-lysine is contained in 1.25 g of L-lysine HCl.
Example foods containing significant proportions of lysine
Food Lysine (% of protein)
Fish 9.19%
Beef, ground, 90% lean/10% fat, cooked 8.31%
Chicken, roasting, meat and skin, cooked, roasted 8.11%
Azuki bean (adzuki beans), mature seeds, raw 7.53%
Milk, non-fat 7.48%
Soybean, mature seeds, raw 7.42%
Egg, whole, raw 7.27%
Pea, split, mature seeds, raw 7.22%
Lentil, pink, raw 6.97%
Kidney bean, mature seeds, raw 6.87%
Chickpea, (garbanzo beans, Bengal gram), mature seeds, raw 6.69%
Navy bean, mature seeds, raw 5.73%

Biological roles
The most common role for lysine is proteinogenesis. Lysine frequently plays an important role in protein
structure. Since its side chain contains a positively charged group on one end and a long hydrophobic
carbon tail close to the backbone, lysine is considered somewhat amphipathic. For this reason, lysine can
be found buried as well as more commonly in solvent channels and on the exterior of proteins, where it
can interact with the aqueous environment.[61] Lysine can also contribute to protein stability as its ε-
amino group often participates in hydrogen bonding, salt bridges and covalent interactions to form a
Schiff base.[61][62][63][64]

A second major role of lysine is in epigenetic regulation by means of histone modification.[65][66] There
are several types of covalent histone modifications, which commonly involve lysine residues found in the
protruding tail of histones. Modifications often include the addition or removal of an acetyl (-CH3CO)
forming acetyllysine or reverting to lysine, up to three methyl (‑CH3), ubiquitin or a sumo protein
group.[65][67][68][69][70] The various modifications have downstream effects on gene regulation, in which
genes can be activated or repressed.

Lysine has also been implicated to play a key role in other biological processes including; structural
proteins of connective tissues, calcium homeostasis, and fatty acid metabolism.[71][72][73] Lysine has
been shown to be involved in the crosslinking between the three helical polypeptides in collagen,
resulting in its stability and tensile strength.[71][74] This mechanism is akin to the role of lysine in
bacterial cell walls, in which lysine (and meso-diaminopimelate) are critical to the formation of
crosslinks, and therefore, stability of the cell wall.[75] This concept has previously been explored as a
means to circumvent the unwanted release of potentially pathogenic genetically modified bacteria. It was
proposed that an auxotrophic strain of Escherichia coli (X1776) could be used for all genetic
modification practices, as the strain is unable to survive without the supplementation of DAP, and thus,
cannot live outside of a laboratory environment.[76] Lysine has also been proposed to be involved in
calcium intestinal absorption and renal retention, and thus, may play a role in calcium homeostasis.[72]
Finally, lysine has been shown to be a precursor for carnitine, which transports fatty acids to the
mitochondria, where they can be oxidised for the release of energy.[73][77] Carnitine is synthesised from
trimethyllysine, which is a product of the degradation of certain proteins, as such lysine must first be
incorporated into proteins and be methylated prior to being converted to carnitine.[73] However, in
mammals the primary source of carnitine is through dietary sources, rather than through lysine
conversion.[73]

In opsins like rhodopsin and the visual opsins (encoded by the genes OPN1SW, OPN1MW, and
OPN1LW), retinaldehyde forms a Schiff base with a conserved lysine residue, and interaction of light
with the retinylidene group causes signal transduction in color vision (See visual cycle for details).

Disputed roles
There has been a long discussion that lysine, when administered intravenously or orally, can significantly
increase the release of growth hormones.[78] This has led to athletes using lysine as a means of promoting
muscle growth while training, however, no significant evidence to support this application of lysine has
been found to date.[78][79]

Because herpes simplex virus (HSV) proteins are richer in arginine and poorer in lysine than the cells
they infect, lysine supplements have been tried as a treatment. Since the two amino acids are taken up in
the intestine, reclaimed in the kidney, and moved into cells by the same amino acid transporters, an
abundance of lysine would, in theory, limit the amount of arginine available for viral replication.[80]
Clinical studies do not provide good evidence for effectiveness as a prophylactic or in the treatment for
HSV outbreaks.[81][82] In response to product claims that lysine could improve immune responses to
HSV, a review by the European Food Safety Authority found no evidence of a cause-effect relationship.
The same review, published in 2011, found no evidence to support claims that lysine could lower
cholesterol, increase appetite, contribute to protein synthesis in any role other than as an ordinary
nutrient, or increase calcium absorption or retention.[83]

Roles in disease
Diseases related to lysine are a result of the downstream processing of lysine, i.e. the incorporation into
proteins or modification into alternative biomolecules. The role of lysine in collagen has been outlined
above, however, a lack of lysine and hydroxylysine involved in the crosslinking of collagen peptides has
been linked to a disease state of the connective tissue.[84] As carnitine is a key lysine-derived metabolite
involved in fatty acid metabolism, a substandard diet lacking sufficient carnitine and lysine can lead to
decreased carnitine levels, which can have significant cascading effects on an individual's health.[77][85]
Lysine has also been shown to play a role in anaemia, as lysine is suspected to have an effect on the
uptake of iron and, subsequently, the concentration of ferritin in blood plasma.[86] However, the exact
mechanism of action is yet to be elucidated.[86] Most commonly, lysine deficiency is seen in non-western
societies and manifests as protein-energy malnutrition, which has profound and systemic effects on the
health of the individual.[87][88] There is also a hereditary genetic disease that involves mutations in the
enzymes responsible for lysine catabolism, namely the bifunctional AASS enzyme of the saccharopine
pathway.[89] Due to a lack of lysine catabolism, the amino acid accumulates in plasma and patients
develop hyperlysinaemia, which can present as asymptomatic to severe neurological disabilities,
including epilepsy, ataxia, spasticity, and psychomotor impairment.[89][90] The clinical significance of
hyperlysinemia is the subject of debate in the field with some studies finding no correlation between
physical or mental disabilities and hyperlysinemia.[91] In addition to this, mutations in genes related to
lysine metabolism have been implicated in several disease states, including pyridoxine-dependent
epilepsia (ALDH7A1 gene), α-ketoadipic and α-aminoadipic aciduria (DHTKD1 gene), and glutaric
aciduria type 1 (GCDH gene).[41][92][93][94][95]

Hyperlysinuria is marked by high amounts of lysine in the urine.[96] It is often due to a metabolic disease
in which a protein involved in the breakdown of lysine is non functional due to a genetic mutation.[97] It
may also occur due to a failure of renal tubular transport.[97]

Use of lysine in animal feed


Lysine production for animal feed is a major global industry, reaching in 2009 almost 700,000 tonnes for
a market value of over €1.22 billion.[98] Lysine is an important additive to animal feed because it is a
limiting amino acid when optimizing the growth of certain animals such as pigs and chickens for the
production of meat. Lysine supplementation allows for the use of lower-cost plant protein (maize, for
instance, rather than soy) while maintaining high growth rates, and limiting the pollution from nitrogen
excretion.[99] In turn, however, phosphate pollution is a major environmental cost when corn is used as
feed for poultry and swine.[100]

Lysine is industrially produced by microbial fermentation, from a base mainly of sugar. Genetic
engineering research is actively pursuing bacterial strains to improve the efficiency of production and
allow lysine to be made from other substrates.[98]

In popular culture
The 1993 film Jurassic Park (based on the 1990 Michael Crichton novel of the same name) features
dinosaurs that were genetically altered so that they could not produce lysine, an example of engineered
auxotrophy.[101] This was known as the "lysine contingency" and was supposed to prevent the cloned
dinosaurs from surviving outside the park, forcing them to be dependent on lysine supplements provided
by the park's veterinary staff. In reality, no animals are capable of producing lysine (it is an essential
amino acid).[102]

In 1996, lysine became the focus of a price-fixing case, the largest in United States history. The Archer
Daniels Midland Company paid a fine of US$100 million, and three of its executives were convicted and
served prison time. Also found guilty in the price-fixing case were two Japanese firms (Ajinomoto,
Kyowa Hakko) and a South Korean firm (Sewon).[103] Secret video recordings of the conspirators fixing
lysine's price can be found online or by requesting the video from the U.S. Department of Justice,
Antitrust Division. This case served as the basis of the movie The Informant!, and a book of the same
title.[104]

References
1. "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). Nomenclature and
symbolism for amino acids and peptides. Recommendations 1983" (https://www.ncbi.nlm.ni
h.gov/pmc/articles/PMC1153490). Biochemical Journal. 219 (2): 345–373. 15 April 1984.
doi:10.1042/bj2190345 (https://doi.org/10.1042%2Fbj2190345). PMC 1153490 (https://ww
w.ncbi.nlm.nih.gov/pmc/articles/PMC1153490). PMID 6743224 (https://www.ncbi.nlm.nih.go
v/pubmed/6743224).
2. Lysine. (http://www.biology.arizona.edu/biochemistry/problem_sets/aa/Lysine.html) The
Biology Project, Department of Biochemistry and Molecular Biophysics, University of
Arizona.
3. Drechsel E (1889). "Zur Kenntniss der Spaltungsprodukte des Caseïns" (https://babel.hathit
rust.org/cgi/pt?id=osu.32435060196680;view=1up;seq=441) [[Contribution] to [our]
knowledge of the cleavage products of casein]. Journal für Praktische Chemie. 2nd series
(in German). 39: 425–429. doi:10.1002/prac.18890390135 (https://doi.org/10.1002%2Fprac.
18890390135). On p. 428, Drechsel presented an empirical formula for the chloroplatinate
salt of lysine – C8H16N2O2Cl2•PtCl4 + H2O – but he later admitted that this formula was
wrong because the salt's crystals contained ethanol instead of water. See: Drechsel E
(1891). "Der Abbau der Eiweissstoffe" [The disassembly of proteins]. Archiv für Anatomie
und Physiologie (in German): 248–278.; Drechsel E. "Zur Kenntniss der Spaltungsproducte
des Caseïns" (https://babel.hathitrust.org/cgi/pt?id=coo.31924056294253;view=1up;seq=26
8) [Contribution] to [our] knowledge of the cleavage products of casein] (in German): 254–
260. "From p. 256:] " … die darin enthaltene Base hat die Formel C6H14N2O2. Der
anfängliche Irrthum ist dadurch veranlasst worden, dass das Chloroplatinat nicht, wie
angenommen ward, Krystallwasser, sondern Krystallalkohol enthält, … " ( … the base
[that's] contained therein has the [empirical] formula C6H14N2O2. The initial error was
caused by the chloroplatinate containing not water in the crystal (as was assumed), but
ethanol … )"
4. Drechsel E (1891). "Der Abbau der Eiweissstoffe" [The disassembly of proteins]. Archiv für
Anatomie und Physiologie (in German): 248–278.; Fischer E (1891). "Ueber neue
Spaltungsproducte des Leimes" (https://babel.hathitrust.org/cgi/pt?id=coo.3192405629425
3;view=1up;seq=281) [On new cleavage products of gelatin] (in German): 465–469. "From
p. 469:] " … die Base C6H14N2O2, welche mit dem Namen Lysin bezeichnet werden mag,
… " ( … the base C6H14N2O2, which may be designated with the name "lysine", … ) [Note:
Ernst Fischer was a graduate student of Drechsel.]"
5. Fischer E, Weigert F (1902). "Synthese der α,ε – Diaminocapronsäure (Inactives Lysin)" (htt
ps://babel.hathitrust.org/cgi/pt?id=hvd.cl1i27;view=1up;seq=1240) [Synthesis of α,ε-
diaminohexanoic acid ([optically] inactive lysine)]. Berichte der Deutschen Chemischen
Gesellschaft (in German). 35 (3): 3772–3778. doi:10.1002/cber.190203503211 (https://doi.o
rg/10.1002%2Fcber.190203503211).
6. Hudson AO, Bless C, Macedo P, Chatterjee SP, Singh BK, Gilvarg C, Leustek T (January
2005). "Biosynthesis of lysine in plants: evidence for a variant of the known bacterial
pathways". Biochimica et Biophysica Acta. 1721 (1–3): 27–36.
doi:10.1016/j.bbagen.2004.09.008 (https://doi.org/10.1016%2Fj.bbagen.2004.09.008).
PMID 15652176 (https://www.ncbi.nlm.nih.gov/pubmed/15652176).
7. Velasco AM, Leguina JI, Lazcano A (October 2002). "Molecular evolution of the lysine
biosynthetic pathways". Journal of Molecular Evolution. 55 (4): 445–59.
doi:10.1007/s00239-002-2340-2 (https://doi.org/10.1007%2Fs00239-002-2340-2).
PMID 12355264 (https://www.ncbi.nlm.nih.gov/pubmed/12355264).
8. Miyazaki T, Miyazaki J, Yamane H, Nishiyama M (July 2004). "alpha-Aminoadipate
aminotransferase from an extremely thermophilic bacterium, Thermus thermophilus".
Microbiology. 150 (Pt 7): 2327–34. doi:10.1099/mic.0.27037-0 (https://doi.org/10.1099%2F
mic.0.27037-0). PMID 15256574 (https://www.ncbi.nlm.nih.gov/pubmed/15256574).
9. Xu H, Andi B, Qian J, West AH, Cook PF (2006). "The alpha-aminoadipate pathway for
lysine biosynthesis in fungi". Cell Biochemistry and Biophysics. 46 (1): 43–64.
doi:10.1385/CBB:46:1:43 (https://doi.org/10.1385%2FCBB%3A46%3A1%3A43).
PMID 16943623 (https://www.ncbi.nlm.nih.gov/pubmed/16943623).
10. Atkinson SC, Dogovski C, Downton MT, Czabotar PE, Dobson RC, Gerrard JA, Wagner J,
Perugini MA (March 2013). "Structural, kinetic and computational investigation of Vitis
vinifera DHDPS reveals new insight into the mechanism of lysine-mediated allosteric
inhibition". Plant Molecular Biology. 81 (4–5): 431–46. doi:10.1007/s11103-013-0014-7 (http
s://doi.org/10.1007%2Fs11103-013-0014-7). PMID 23354837 (https://www.ncbi.nlm.nih.gov/
pubmed/23354837).
11. Griffin MD, Billakanti JM, Wason A, Keller S, Mertens HD, Atkinson SC, Dobson RC,
Perugini MA, Gerrard JA, Pearce FG (2012). "Characterisation of the first enzymes
committed to lysine biosynthesis in Arabidopsis thaliana" (https://www.ncbi.nlm.nih.gov/pmc/
articles/PMC3390394). PLOS ONE. 7 (7): e40318. doi:10.1371/journal.pone.0040318 (http
s://doi.org/10.1371%2Fjournal.pone.0040318). PMC 3390394 (https://www.ncbi.nlm.nih.go
v/pmc/articles/PMC3390394). PMID 22792278 (https://www.ncbi.nlm.nih.gov/pubmed/2279
2278).
12. Soares da Costa TP, Muscroft-Taylor AC, Dobson RC, Devenish SR, Jameson GB, Gerrard
JA (July 2010). "How essential is the 'essential' active-site lysine in dihydrodipicolinate
synthase?". Biochimie. 92 (7): 837–45. doi:10.1016/j.biochi.2010.03.004 (https://doi.org/10.
1016%2Fj.biochi.2010.03.004). PMID 20353808 (https://www.ncbi.nlm.nih.gov/pubmed/203
53808).
13. Soares da Costa TP, Christensen JB, Desbois S, Gordon SE, Gupta R, Hogan CJ, Nelson
TG, Downton MT, Gardhi CK, Abbott BM, Wagner J, Panjikar S, Perugini MA (2015).
"Quaternary Structure Analyses of an Essential Oligomeric Enzyme". Analytical
Ultracentrifugation. Methods in Enzymology. 562. pp. 205–23.
doi:10.1016/bs.mie.2015.06.020 (https://doi.org/10.1016%2Fbs.mie.2015.06.020).
ISBN 9780128029084. PMID 26412653 (https://www.ncbi.nlm.nih.gov/pubmed/26412653).
14. Muscroft-Taylor AC, Soares da Costa TP, Gerrard JA (March 2010). "New insights into the
mechanism of dihydrodipicolinate synthase using isothermal titration calorimetry".
Biochimie. 92 (3): 254–62. doi:10.1016/j.biochi.2009.12.004 (https://doi.org/10.1016%2Fj.bi
ochi.2009.12.004). PMID 20025926 (https://www.ncbi.nlm.nih.gov/pubmed/20025926).
15. Christensen JB, Soares da Costa TP, Faou P, Pearce FG, Panjikar S, Perugini MA
(November 2016). "Structure and Function of Cyanobacterial DHDPS and DHDPR" (https://
www.ncbi.nlm.nih.gov/pmc/articles/PMC5109050). Scientific Reports. 6 (1): 37111.
doi:10.1038/srep37111 (https://doi.org/10.1038%2Fsrep37111). PMC 5109050 (https://ww
w.ncbi.nlm.nih.gov/pmc/articles/PMC5109050). PMID 27845445 (https://www.ncbi.nlm.nih.g
ov/pubmed/27845445).
16. McCoy AJ, Adams NE, Hudson AO, Gilvarg C, Leustek T, Maurelli AT (November 2006).
"L,L-diaminopimelate aminotransferase, a trans-kingdom enzyme shared by Chlamydia and
plants for synthesis of diaminopimelate/lysine" (https://www.ncbi.nlm.nih.gov/pmc/articles/P
MC1693846). Proceedings of the National Academy of Sciences of the United States of
America. 103 (47): 17909–14. doi:10.1073/pnas.0608643103 (https://doi.org/10.1073%2Fp
nas.0608643103). PMC 1693846 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1693846).
PMID 17093042 (https://www.ncbi.nlm.nih.gov/pubmed/17093042).
17. Hudson AO, Gilvarg C, Leustek T (May 2008). "Biochemical and phylogenetic
characterization of a novel diaminopimelate biosynthesis pathway in prokaryotes identifies a
diverged form of LL-diaminopimelate aminotransferase" (https://www.ncbi.nlm.nih.gov/pmc/
articles/PMC2347407). Journal of Bacteriology. 190 (9): 3256–63. doi:10.1128/jb.01381-07
(https://doi.org/10.1128%2Fjb.01381-07). PMC 2347407 (https://www.ncbi.nlm.nih.gov/pmc/
articles/PMC2347407). PMID 18310350 (https://www.ncbi.nlm.nih.gov/pubmed/18310350).
18. Peverelli MG, Perugini MA (August 2015). "An optimized coupled assay for quantifying
diaminopimelate decarboxylase activity". Biochimie. 115: 78–85.
doi:10.1016/j.biochi.2015.05.004 (https://doi.org/10.1016%2Fj.biochi.2015.05.004).
PMID 25986217 (https://www.ncbi.nlm.nih.gov/pubmed/25986217).
19. Soares da Costa TP, Desbois S, Dogovski C, Gorman MA, Ketaren NE, Paxman JJ,
Siddiqui T, Zammit LM, Abbott BM, Robins-Browne RM, Parker MW, Jameson GB, Hall NE,
Panjikar S, Perugini MA (August 2016). "Structural Determinants Defining the Allosteric
Inhibition of an Essential Antibiotic Target". Structure. 24 (8): 1282–1291.
doi:10.1016/j.str.2016.05.019 (https://doi.org/10.1016%2Fj.str.2016.05.019).
PMID 27427481 (https://www.ncbi.nlm.nih.gov/pubmed/27427481).
20. Jander G, Joshi V (1 January 2009). "Aspartate-Derived Amino Acid Biosynthesis in
Arabidopsis thaliana" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3243338). The
Arabidopsis Book. 7: e0121. doi:10.1199/tab.0121 (https://doi.org/10.1199%2Ftab.0121).
PMC 3243338 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3243338). PMID 22303247
(https://www.ncbi.nlm.nih.gov/pubmed/22303247).
21. Andi B, West AH, Cook PF (September 2004). "Kinetic mechanism of histidine-tagged
homocitrate synthase from Saccharomyces cerevisiae". Biochemistry. 43 (37): 11790–5.
doi:10.1021/bi048766p (https://doi.org/10.1021%2Fbi048766p). PMID 15362863 (https://ww
w.ncbi.nlm.nih.gov/pubmed/15362863).
22. Bhattacharjee JK (1985). "alpha-Aminoadipate pathway for the biosynthesis of lysine in
lower eukaryotes". Critical Reviews in Microbiology. 12 (2): 131–51.
doi:10.3109/10408418509104427 (https://doi.org/10.3109%2F10408418509104427).
PMID 3928261 (https://www.ncbi.nlm.nih.gov/pubmed/3928261).
23. Bhattacharjee JK, Strassman M (May 1967). "Accumulation of tricarboxylic acids related to
lysine biosynthesis in a yeast mutant". The Journal of Biological Chemistry. 242 (10): 2542–
6. PMID 6026248 (https://www.ncbi.nlm.nih.gov/pubmed/6026248).
24. Gaillardin CM, Ribet AM, Heslot H (November 1982). "Wild-type and mutant forms of
homoisocitric dehydrogenase in the yeast Saccharomycopsis lipolytica". European Journal
of Biochemistry. 128 (2–3): 489–94. doi:10.1111/j.1432-1033.1982.tb06991.x (https://doi.or
g/10.1111%2Fj.1432-1033.1982.tb06991.x). PMID 6759120 (https://www.ncbi.nlm.nih.gov/p
ubmed/6759120).
25. Jaklitsch WM, Kubicek CP (July 1990). "Homocitrate synthase from Penicillium
chrysogenum. Localization, purification of the cytosolic isoenzyme, and sensitivity to lysine"
(https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1131560). The Biochemical Journal. 269
(1): 247–53. doi:10.1042/bj2690247 (https://doi.org/10.1042%2Fbj2690247). PMC 1131560
(https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1131560). PMID 2115771 (https://www.ncbi.
nlm.nih.gov/pubmed/2115771).
26. Ye ZH, Bhattacharjee JK (December 1988). "Lysine biosynthesis pathway and biochemical
blocks of lysine auxotrophs of Schizosaccharomyces pombe" (https://www.ncbi.nlm.nih.gov/
pmc/articles/PMC211717). Journal of Bacteriology. 170 (12): 5968–70.
doi:10.1128/jb.170.12.5968-5970.1988 (https://doi.org/10.1128%2Fjb.170.12.5968-5970.19
88). PMC 211717 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC211717). PMID 3142867
(https://www.ncbi.nlm.nih.gov/pubmed/3142867).
27. Kobashi N, Nishiyama M, Tanokura M (March 1999). "Aspartate kinase-independent lysine
synthesis in an extremely thermophilic bacterium, Thermus thermophilus: lysine is
synthesized via alpha-aminoadipic acid not via diaminopimelic acid" (https://www.ncbi.nlm.ni
h.gov/pmc/articles/PMC93567). Journal of Bacteriology. 181 (6): 1713–8. PMC 93567 (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC93567). PMID 10074061 (https://www.ncbi.nlm.n
ih.gov/pubmed/10074061).
28. Kosuge T, Hoshino T (1999). "The alpha-aminoadipate pathway for lysine biosynthesis is
widely distributed among Thermus strains". Journal of Bioscience and Bioengineering. 88
(6): 672–5. doi:10.1016/S1389-1723(00)87099-1 (https://doi.org/10.1016%2FS1389-1723%
2800%2987099-1). PMID 16232683 (https://www.ncbi.nlm.nih.gov/pubmed/16232683).
29. Nishida H, Nishiyama M, Kobashi N, Kosuge T, Hoshino T, Yamane H (December 1999). "A
prokaryotic gene cluster involved in synthesis of lysine through the amino adipate pathway:
a key to the evolution of amino acid biosynthesis". Genome Research. 9 (12): 1175–83.
doi:10.1101/gr.9.12.1175 (https://doi.org/10.1101%2Fgr.9.12.1175). PMID 10613839 (http
s://www.ncbi.nlm.nih.gov/pubmed/10613839).
30. Nishida H, Nishiyama M (September 2000). "What is characteristic of fungal lysine
synthesis through the alpha-aminoadipate pathway?". Journal of Molecular Evolution. 51
(3): 299–302. doi:10.1007/s002390010091 (https://doi.org/10.1007%2Fs002390010091).
PMID 11029074 (https://www.ncbi.nlm.nih.gov/pubmed/11029074).
31. Zabriskie TM, Jackson MD (February 2000). "Lysine biosynthesis and metabolism in fungi".
Natural Product Reports. 17 (1): 85–97. doi:10.1039/a801345d (https://doi.org/10.1039%2F
a801345d). PMID 10714900 (https://www.ncbi.nlm.nih.gov/pubmed/10714900).
32. Zhu X, Galili G (May 2004). "Lysine metabolism is concurrently regulated by synthesis and
catabolism in both reproductive and vegetative tissues" (https://www.ncbi.nlm.nih.gov/pmc/a
rticles/PMC429340). Plant Physiology. 135 (1): 129–36. doi:10.1104/pp.103.037168 (https://
doi.org/10.1104%2Fpp.103.037168). PMC 429340 (https://www.ncbi.nlm.nih.gov/pmc/articl
es/PMC429340). PMID 15122025 (https://www.ncbi.nlm.nih.gov/pubmed/15122025).
33. Tomé D, Bos C (June 2007). "Lysine requirement through the human life cycle". The Journal
of Nutrition. 137 (6 Suppl 2): 1642S–1645S. doi:10.1093/jn/137.6.1642S (https://doi.org/10.
1093%2Fjn%2F137.6.1642S). PMID 17513440 (https://www.ncbi.nlm.nih.gov/pubmed/1751
3440).
34. Blemings KP, Crenshaw TD, Swick RW, Benevenga NJ (August 1994). "Lysine-alpha-
ketoglutarate reductase and saccharopine dehydrogenase are located only in the
mitochondrial matrix in rat liver". The Journal of Nutrition. 124 (8): 1215–21.
doi:10.1093/jn/124.8.1215 (https://doi.org/10.1093%2Fjn%2F124.8.1215). PMID 8064371
(https://www.ncbi.nlm.nih.gov/pubmed/8064371).
35. Galili G, Tang G, Zhu X, Gakiere B (June 2001). "Lysine catabolism: a stress and
development super-regulated metabolic pathway". Current Opinion in Plant Biology. 4 (3):
261–6. doi:10.1016/s1369-5266(00)00170-9 (https://doi.org/10.1016%2Fs1369-5266%280
0%2900170-9). PMID 11312138 (https://www.ncbi.nlm.nih.gov/pubmed/11312138).
36. Arruda P, Kemper EL, Papes F, Leite A (August 2000). "Regulation of lysine catabolism in
higher plants". Trends in Plant Science. 5 (8): 324–30. doi:10.1016/s1360-1385(00)01688-5
(https://doi.org/10.1016%2Fs1360-1385%2800%2901688-5). PMID 10908876 (https://www.
ncbi.nlm.nih.gov/pubmed/10908876).
37. Sacksteder KA, Biery BJ, Morrell JC, Goodman BK, Geisbrecht BV, Cox RP, Gould SJ,
Geraghty MT (June 2000). "Identification of the alpha-aminoadipic semialdehyde synthase
gene, which is defective in familial hyperlysinemia" (https://www.ncbi.nlm.nih.gov/pmc/article
s/PMC1378037). American Journal of Human Genetics. 66 (6): 1736–43.
doi:10.1086/302919 (https://doi.org/10.1086%2F302919). PMC 1378037 (https://www.ncbi.
nlm.nih.gov/pmc/articles/PMC1378037). PMID 10775527 (https://www.ncbi.nlm.nih.gov/pub
med/10775527).
38. Zhu X, Tang G, Galili G (December 2002). "The activity of the Arabidopsis bifunctional
lysine-ketoglutarate reductase/saccharopine dehydrogenase enzyme of lysine catabolism is
regulated by functional interaction between its two enzyme domains". The Journal of
Biological Chemistry. 277 (51): 49655–61. doi:10.1074/jbc.m205466200 (https://doi.org/10.1
074%2Fjbc.m205466200). PMID 12393892 (https://www.ncbi.nlm.nih.gov/pubmed/1239389
2).
39. Kiyota E, Pena IA, Arruda P (November 2015). "The saccharopine pathway in seed
development and stress response of maize". Plant, Cell & Environment. 38 (11): 2450–61.
doi:10.1111/pce.12563 (https://doi.org/10.1111%2Fpce.12563). PMID 25929294 (https://ww
w.ncbi.nlm.nih.gov/pubmed/25929294).
40. Serrano GC, Rezende e Silva Figueira T, Kiyota E, Zanata N, Arruda P (March 2012).
"Lysine degradation through the saccharopine pathway in bacteria: LKR and SDH in
bacteria and its relationship to the plant and animal enzymes". FEBS Letters. 586 (6): 905–
11. doi:10.1016/j.febslet.2012.02.023 (https://doi.org/10.1016%2Fj.febslet.2012.02.023).
PMID 22449979 (https://www.ncbi.nlm.nih.gov/pubmed/22449979).
41. Danhauser K, Sauer SW, Haack TB, Wieland T, Staufner C, Graf E, Zschocke J, Strom TM,
Traub T, Okun JG, Meitinger T, Hoffmann GF, Prokisch H, Kölker S (December 2012).
"DHTKD1 mutations cause 2-aminoadipic and 2-oxoadipic aciduria" (https://www.ncbi.nlm.ni
h.gov/pmc/articles/PMC3516599). American Journal of Human Genetics. 91 (6): 1082–7.
doi:10.1016/j.ajhg.2012.10.006 (https://doi.org/10.1016%2Fj.ajhg.2012.10.006).
PMC 3516599 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3516599). PMID 23141293
(https://www.ncbi.nlm.nih.gov/pubmed/23141293).
42. Sauer SW, Opp S, Hoffmann GF, Koeller DM, Okun JG, Kölker S (January 2011).
"Therapeutic modulation of cerebral L-lysine metabolism in a mouse model for glutaric
aciduria type I". Brain. 134 (Pt 1): 157–70. doi:10.1093/brain/awq269 (https://doi.org/10.109
3%2Fbrain%2Fawq269). PMID 20923787
(https://www.ncbi.nlm.nih.gov/pubmed/20923787).
43. Goncalves RL, Bunik VI, Brand MD (February 2016). "Production of superoxide/hydrogen
peroxide by the mitochondrial 2-oxoadipate dehydrogenase complex". Free Radical Biology
& Medicine. 91: 247–55. doi:10.1016/j.freeradbiomed.2015.12.020 (https://doi.org/10.101
6%2Fj.freeradbiomed.2015.12.020). PMID 26708453 (https://www.ncbi.nlm.nih.gov/pubme
d/26708453).
44. Goh DL, Patel A, Thomas GH, Salomons GS, Schor DS, Jakobs C, Geraghty MT (July
2002). "Characterization of the human gene encoding alpha-aminoadipate
aminotransferase (AADAT)". Molecular Genetics and Metabolism. 76 (3): 172–80.
doi:10.1016/s1096-7192(02)00037-9 (https://doi.org/10.1016%2Fs1096-7192%2802%2900
037-9). PMID 12126930 (https://www.ncbi.nlm.nih.gov/pubmed/12126930).
45. Härtel U, Eckel E, Koch J, Fuchs G, Linder D, Buckel W (1 February 1993). "Purification of
glutaryl-CoA dehydrogenase from Pseudomonas sp., an enzyme involved in the anaerobic
degradation of benzoate". Archives of Microbiology. 159 (2): 174–81.
doi:10.1007/bf00250279 (https://doi.org/10.1007%2Fbf00250279). PMID 8439237 (https://w
ww.ncbi.nlm.nih.gov/pubmed/8439237).
46. Sauer SW (October 2007). "Biochemistry and bioenergetics of glutaryl-CoA dehydrogenase
deficiency". Journal of Inherited Metabolic Disease. 30 (5): 673–80. doi:10.1007/s10545-
007-0678-8 (https://doi.org/10.1007%2Fs10545-007-0678-8). PMID 17879145 (https://www.
ncbi.nlm.nih.gov/pubmed/17879145).
47. Nelson DL, Cox MM, Lehninger AL (2013). Lehninger principles of biochemistry (6th ed.).
New York: W.H. Freeman and Company. ISBN 978-1-4641-0962-1. OCLC 824794893 (http
s://www.worldcat.org/oclc/824794893).
48. Galili G, Amir R (February 2013). "Fortifying plants with the essential amino acids lysine and
methionine to improve nutritional quality". Plant Biotechnology Journal. 11 (2): 211–22.
doi:10.1111/pbi.12025 (https://doi.org/10.1111%2Fpbi.12025). PMID 23279001 (https://ww
w.ncbi.nlm.nih.gov/pubmed/23279001).
49. Wang G, Xu M, Wang W, Galili G (June 2017). "Fortifying Horticultural Crops with Essential
Amino Acids: A Review" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5486127).
International Journal of Molecular Sciences. 18 (6): 1306. doi:10.3390/ijms18061306 (http
s://doi.org/10.3390%2Fijms18061306). PMC 5486127 (https://www.ncbi.nlm.nih.gov/pmc/art
icles/PMC5486127). PMID 28629176 (https://www.ncbi.nlm.nih.gov/pubmed/28629176).
50. Angelovici R, Fait A, Fernie AR, Galili G (January 2011). "A seed high-lysine trait is
negatively associated with the TCA cycle and slows down Arabidopsis seed germination".
The New Phytologist. 189 (1): 148–59. doi:10.1111/j.1469-8137.2010.03478.x (https://doi.or
g/10.1111%2Fj.1469-8137.2010.03478.x). PMID 20946418 (https://www.ncbi.nlm.nih.gov/p
ubmed/20946418).
51. Edelman M, Colt M (2016). "Nutrient Value of Leaf vs. Seed" (https://www.ncbi.nlm.nih.gov/
pmc/articles/PMC4954856). Frontiers in Chemistry. 4: 32. doi:10.3389/fchem.2016.00032 (h
ttps://doi.org/10.3389%2Ffchem.2016.00032). PMC 4954856 (https://www.ncbi.nlm.nih.gov/
pmc/articles/PMC4954856). PMID 27493937 (https://www.ncbi.nlm.nih.gov/pubmed/274939
37).
52. Jiang SY, Ma A, Xie L, Ramachandran S (September 2016). "Improving protein content and
quality by over-expressing artificially synthetic fusion proteins with high lysine and threonine
constituent in rice plants" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5039639).
Scientific Reports. 6 (1): 34427. doi:10.1038/srep34427 (https://doi.org/10.1038%2Fsrep34
427). PMC 5039639 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5039639).
PMID 27677708 (https://www.ncbi.nlm.nih.gov/pubmed/27677708).
53. Shewry PR (November 2007). "Improving the protein content and composition of cereal
grain". Journal of Cereal Science. 46 (3): 239–250. doi:10.1016/j.jcs.2007.06.006 (https://do
i.org/10.1016%2Fj.jcs.2007.06.006).
54. Prasanna B, Vasal SK, Kassahun B, Singh NN (2001). "Quality protein maize". Current
Science. 81 (10): 1308–1319. JSTOR 24105845 (https://www.jstor.org/stable/24105845).
55. Kircher M, Pfefferle W (April 2001). "The fermentative production of L-lysine as an animal
feed additive". Chemosphere. 43 (1): 27–31. doi:10.1016/s0045-6535(00)00320-9 (https://d
oi.org/10.1016%2Fs0045-6535%2800%2900320-9). PMID 11233822 (https://www.ncbi.nlm.
nih.gov/pubmed/11233822).
56. Junior L, Alberto L, Letti GV, Soccol CR, Junior L, Alberto L, Letti GV, Soccol CR (2016).
"Development of an L-Lysine Enriched Bran for Animal Nutrition via Submerged
Fermentation by Corynebacterium glutamicum using Agroindustrial Substrates". Brazilian
Archives of Biology and Technology. 59. doi:10.1590/1678-4324-2016150519 (https://doi.or
g/10.1590%2F1678-4324-2016150519). ISSN 1516-8913 (https://www.worldcat.org/issn/15
16-8913).
57. University of Maryland Medical Center. "Lysine" (http://www.umm.edu/altmed/articles/lysine-
000312.htm). Retrieved 30 December 2009.
58. Young VR, Pellett PL (1994). "Plant proteins in relation to human protein and amino acid
nutrition"
(https://semanticscholar.org/paper/af389c20178e50ef0187ff3fe94cb47a401e9a8b).
American Journal of Clinical Nutrition. 59 (5&nbsp, Suppl): 1203S–1212S.
doi:10.1093/ajcn/59.5.1203s (https://doi.org/10.1093%2Fajcn%2F59.5.1203s).
PMID 8172124 (https://www.ncbi.nlm.nih.gov/pubmed/8172124).
59. Institute of Medicine of the National Academies. "Dietary Reference Intakes for
Macronutrients" (https://www.nap.edu/read/10490/chapter/12). p. 589. Retrieved 29 October
2017.
60. "Dietary Supplement Database: Blend Information (DSBI)" (https://www.cdc.gov/nchs/nhane
s/nhanes1999-2000/DSBI.htm). "L-LYSINE HCL 10000820 80.03% lysine"
61. Betts MJ, Russell RB (2003). Barnes MR, Gray IC (eds.). Bioinformatics for Geneticists.
John Wiley & Sons, Ltd. pp. 289–316. doi:10.1002/0470867302.ch14 (https://doi.org/10.100
2%2F0470867302.ch14). ISBN 978-0-470-86730-3.
62. Blickling S, Renner C, Laber B, Pohlenz HD, Holak TA, Huber R (January 1997). "Reaction
mechanism of Escherichia coli dihydrodipicolinate synthase investigated by X-ray
crystallography and NMR spectroscopy" (https://semanticscholar.org/paper/a874955a99a1d
c0d4947b609d1b3fb49851a5e25). Biochemistry. 36 (1): 24–33. doi:10.1021/bi962272d (htt
ps://doi.org/10.1021%2Fbi962272d). PMID 8993314 (https://www.ncbi.nlm.nih.gov/pubmed/
8993314).
63. Kumar S, Tsai CJ, Nussinov R (March 2000). "Factors enhancing protein thermostability".
Protein Engineering. 13 (3): 179–91. doi:10.1093/protein/13.3.179 (https://doi.org/10.1093%
2Fprotein%2F13.3.179). PMID 10775659
(https://www.ncbi.nlm.nih.gov/pubmed/10775659).
64. Sokalingam S, Raghunathan G, Soundrarajan N, Lee SG (9 July 2012). "A study on the
effect of surface lysine to arginine mutagenesis on protein stability and structure using
green fluorescent protein" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3392243). PLOS
ONE. 7 (7): e40410. doi:10.1371/journal.pone.0040410 (https://doi.org/10.1371%2Fjournal.
pone.0040410). PMC 3392243 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3392243).
PMID 22792305 (https://www.ncbi.nlm.nih.gov/pubmed/22792305).
65. Dambacher S, Hahn M, Schotta G (July 2010). "Epigenetic regulation of development by
histone lysine methylation". Heredity. 105 (1): 24–37. doi:10.1038/hdy.2010.49 (https://doi.o
rg/10.1038%2Fhdy.2010.49). PMID 20442736 (https://www.ncbi.nlm.nih.gov/pubmed/20442
736).
66. Martin C, Zhang Y (November 2005). "The diverse functions of histone lysine methylation".
Nature Reviews. Molecular Cell Biology. 6 (11): 838–49. doi:10.1038/nrm1761 (https://doi.or
g/10.1038%2Fnrm1761). PMID 16261189 (https://www.ncbi.nlm.nih.gov/pubmed/1626118
9).
67. Black JC, Van Rechem C, Whetstine JR (November 2012). "Histone lysine methylation
dynamics: establishment, regulation, and biological impact" (https://www.ncbi.nlm.nih.gov/p
mc/articles/PMC3861058). Molecular Cell. 48 (4): 491–507.
doi:10.1016/j.molcel.2012.11.006 (https://doi.org/10.1016%2Fj.molcel.2012.11.006).
PMC 3861058 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3861058). PMID 23200123
(https://www.ncbi.nlm.nih.gov/pubmed/23200123).
68. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, Olsen JV, Mann M
(August 2009). "Lysine acetylation targets protein complexes and co-regulates major
cellular functions" (https://semanticscholar.org/paper/af946911954f7ec506e0afa6faec466ec
e09b117). Science. 325 (5942): 834–40. doi:10.1126/science.1175371 (https://doi.org/10.11
26%2Fscience.1175371). PMID 19608861 (https://www.ncbi.nlm.nih.gov/pubmed/1960886
1).
69. Shiio Y, Eisenman RN (November 2003). "Histone sumoylation is associated with
transcriptional repression" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC263760).
Proceedings of the National Academy of Sciences of the United States of America. 100
(23): 13225–30. doi:10.1073/pnas.1735528100 (https://doi.org/10.1073%2Fpnas.17355281
00). PMC 263760 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC263760). PMID 14578449
(https://www.ncbi.nlm.nih.gov/pubmed/14578449).
70. Wang H, Wang L, Erdjument-Bromage H, Vidal M, Tempst P, Jones RS, Zhang Y (October
2004). "Role of histone H2A ubiquitination in Polycomb silencing". Nature. 431 (7010): 873–
8. doi:10.1038/nature02985 (https://doi.org/10.1038%2Fnature02985). hdl:10261/73732 (htt
ps://hdl.handle.net/10261%2F73732). PMID 15386022 (https://www.ncbi.nlm.nih.gov/pubm
ed/15386022).
71. Shoulders MD, Raines RT (2009). "Collagen structure and stability" (https://www.ncbi.nlm.ni
h.gov/pmc/articles/PMC2846778). Annual Review of Biochemistry. 78: 929–58.
doi:10.1146/annurev.biochem.77.032207.120833 (https://doi.org/10.1146%2Fannurev.bioch
em.77.032207.120833). PMC 2846778 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC284
6778). PMID 19344236 (https://www.ncbi.nlm.nih.gov/pubmed/19344236).
72. Civitelli R, Villareal DT, Agnusdei D, Nardi P, Avioli LV, Gennari C (1992). "Dietary L-lysine
and calcium metabolism in humans". Nutrition. 8 (6): 400–5. PMID 1486246 (https://www.nc
bi.nlm.nih.gov/pubmed/1486246).
73. Vaz FM, Wanders RJ (February 2002). "Carnitine biosynthesis in mammals" (https://www.nc
bi.nlm.nih.gov/pmc/articles/PMC1222323). The Biochemical Journal. 361 (Pt 3): 417–29.
doi:10.1042/bj3610417 (https://doi.org/10.1042%2Fbj3610417). PMC 1222323 (https://ww
w.ncbi.nlm.nih.gov/pmc/articles/PMC1222323). PMID 11802770 (https://www.ncbi.nlm.nih.g
ov/pubmed/11802770).
74. Yamauchi M, Sricholpech M (25 May 2012). "Lysine post-translational modifications of
collagen" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3499978). Essays in
Biochemistry. 52: 113–33. doi:10.1042/bse0520113 (https://doi.org/10.1042%2Fbse052011
3). PMC 3499978 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3499978).
PMID 22708567 (https://www.ncbi.nlm.nih.gov/pubmed/22708567).
75. Vollmer W, Blanot D, de Pedro MA (March 2008). "Peptidoglycan structure and
architecture". FEMS Microbiology Reviews. 32 (2): 149–67. doi:10.1111/j.1574-
6976.2007.00094.x (https://doi.org/10.1111%2Fj.1574-6976.2007.00094.x).
PMID 18194336 (https://www.ncbi.nlm.nih.gov/pubmed/18194336).
76. Curtiss R (May 1978). "Biological containment and cloning vector transmissibility". The
Journal of Infectious Diseases. 137 (5): 668–75. doi:10.1093/infdis/137.5.668 (https://doi.or
g/10.1093%2Finfdis%2F137.5.668). PMID 351084 (https://www.ncbi.nlm.nih.gov/pubmed/3
51084).
77. Flanagan JL, Simmons PA, Vehige J, Willcox MD, Garrett Q (April 2010). "Role of carnitine
in disease" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2861661). Nutrition &
Metabolism. 7: 30. doi:10.1186/1743-7075-7-30 (https://doi.org/10.1186%2F1743-7075-7-3
0). PMC 2861661 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2861661).
PMID 20398344 (https://www.ncbi.nlm.nih.gov/pubmed/20398344).
78. Chromiak JA, Antonio J (2002). "Use of amino acids as growth hormone-releasing agents
by athletes". Nutrition. 18 (7–8): 657–61. doi:10.1016/s0899-9007(02)00807-9 (https://doi.or
g/10.1016%2Fs0899-9007%2802%2900807-9). PMID 12093449 (https://www.ncbi.nlm.nih.
gov/pubmed/12093449).
79. Corpas E, Blackman MR, Roberson R, Scholfield D, Harman SM (July 1993). "Oral
arginine-lysine does not increase growth hormone or insulin-like growth factor-I in old men".
Journal of Gerontology. 48 (4): M128–33. doi:10.1093/geronj/48.4.M128 (https://doi.org/10.
1093%2Fgeronj%2F48.4.M128). PMID 8315224 (https://www.ncbi.nlm.nih.gov/pubmed/831
5224).
80. Gaby AR (2006). "Natural remedies for Herpes simplex". Altern Med Rev. 11 (2): 93–101.
PMID 16813459 (https://www.ncbi.nlm.nih.gov/pubmed/16813459).
81. Tomblin FA, Lucas KH (2001). "Lysine for management of herpes labialis" (https://www.med
scape.com/viewarticle/406943). Am J Health Syst Pharm. 58 (4): 298–300, 304.
doi:10.1093/ajhp/58.4.298 (https://doi.org/10.1093%2Fajhp%2F58.4.298). PMID 11225166
(https://www.ncbi.nlm.nih.gov/pubmed/11225166).
82. Chi CC, Wang SH, Delamere FM, Wojnarowska F, Peters MC, Kanjirath PP (7 August
2015). "Interventions for prevention of herpes simplex labialis (cold sores on the lips)" (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC6461191). The Cochrane Database of
Systematic Reviews (8): CD010095. doi:10.1002/14651858.CD010095.pub2 (https://doi.or
g/10.1002%2F14651858.CD010095.pub2). PMC 6461191 (https://www.ncbi.nlm.nih.gov/pm
c/articles/PMC6461191). PMID 26252373
(https://www.ncbi.nlm.nih.gov/pubmed/26252373).
83. "Scientific Opinion on the substantiation of health claims related to L-lysine and immune
defence against herpes virus (ID 453), maintenance of normal blood LDL-cholesterol
concentrations (ID 454, 4669), increase in appetite leading to an increase in energ". EFSA
Journal. 9 (4): 2063. 2011. doi:10.2903/j.efsa.2011.2063 (https://doi.org/10.2903%2Fj.efsa.
2011.2063). ISSN 1831-4732 (https://www.worldcat.org/issn/1831-4732).
84. Pinnell SR, Krane SM, Kenzora JE, Glimcher MJ (May 1972). "A heritable disorder of
connective tissue. Hydroxylysine-deficient collagen disease". The New England Journal of
Medicine. 286 (19): 1013–20. doi:10.1056/NEJM197205112861901 (https://doi.org/10.105
6%2FNEJM197205112861901). PMID 5016372 (https://www.ncbi.nlm.nih.gov/pubmed/501
6372).
85. Rudman D, Sewell CW, Ansley JD (September 1977). "Deficiency of carnitine in cachectic
cirrhotic patients" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC372417). The Journal of
Clinical Investigation. 60 (3): 716–23. doi:10.1172/jci108824 (https://doi.org/10.1172%2Fjci1
08824). PMC 372417 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC372417).
PMID 893675 (https://www.ncbi.nlm.nih.gov/pubmed/893675).
86. Rushton DH (July 2002). "Nutritional factors and hair loss" (https://semanticscholar.org/pape
r/8c596dd79d93d5f0721a94b9f6d7b5ce6bebfdb9). Clinical and Experimental Dermatology.
27 (5): 396–404. doi:10.1046/j.1365-2230.2002.01076.x (https://doi.org/10.1046%2Fj.1365-
2230.2002.01076.x). PMID 12190640 (https://www.ncbi.nlm.nih.gov/pubmed/12190640).
87. Emery PW (October 2005). "Metabolic changes in malnutrition". Eye. 19 (10): 1029–34.
doi:10.1038/sj.eye.6701959 (https://doi.org/10.1038%2Fsj.eye.6701959). PMID 16304580
(https://www.ncbi.nlm.nih.gov/pubmed/16304580).
88. Ghosh S, Smriga M, Vuvor F, Suri D, Mohammed H, Armah SM, Scrimshaw NS (October
2010). "Effect of lysine supplementation on health and morbidity in subjects belonging to
poor peri-urban households in Accra, Ghana". The American Journal of Clinical Nutrition. 92
(4): 928–39. doi:10.3945/ajcn.2009.28834 (https://doi.org/10.3945%2Fajcn.2009.28834).
PMID 20720257 (https://www.ncbi.nlm.nih.gov/pubmed/20720257).
89. Houten SM, Te Brinke H, Denis S, Ruiter JP, Knegt AC, de Klerk JB, Augoustides-
Savvopoulou P, Häberle J, Baumgartner MR, Coşkun T, Zschocke J, Sass JO, Poll-The BT,
Wanders RJ, Duran M (April 2013). "Genetic basis of hyperlysinemia" (https://www.ncbi.nlm.
nih.gov/pmc/articles/PMC3626681). Orphanet Journal of Rare Diseases. 8: 57.
doi:10.1186/1750-1172-8-57 (https://doi.org/10.1186%2F1750-1172-8-57). PMC 3626681
(https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3626681). PMID 23570448 (https://www.ncb
i.nlm.nih.gov/pubmed/23570448).
90. Hoffmann GF, Kölker S (2016). Inborn Metabolic Diseases. Springer, Berlin, Heidelberg.
pp. 333–348. doi:10.1007/978-3-662-49771-5_22 (https://doi.org/10.1007%2F978-3-662-49
771-5_22). ISBN 978-3-662-49769-2.
91. Dancis J, Hutzler J, Ampola MG, Shih VE, van Gelderen HH, Kirby LT, Woody NC (May
1983). "The prognosis of hyperlysinemia: an interim report" (https://www.ncbi.nlm.nih.gov/p
mc/articles/PMC1685659). American Journal of Human Genetics. 35 (3): 438–42.
PMC 1685659 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1685659). PMID 6407303 (ht
tps://www.ncbi.nlm.nih.gov/pubmed/6407303).
92. Mills PB, Struys E, Jakobs C, Plecko B, Baxter P, Baumgartner M, Willemsen MA, Omran
H, Tacke U, Uhlenberg B, Weschke B, Clayton PT (March 2006). "Mutations in antiquitin in
individuals with pyridoxine-dependent seizures". Nature Medicine. 12 (3): 307–9.
doi:10.1038/nm1366 (https://doi.org/10.1038%2Fnm1366). PMID 16491085 (https://www.nc
bi.nlm.nih.gov/pubmed/16491085).
93. Mills PB, Footitt EJ, Mills KA, Tuschl K, Aylett S, Varadkar S, Hemingway C, Marlow N,
Rennie J, Baxter P, Dulac O, Nabbout R, Craigen WJ, Schmitt B, Feillet F, Christensen E,
De Lonlay P, Pike MG, Hughes MI, Struys EA, Jakobs C, Zuberi SM, Clayton PT (July
2010). "Genotypic and phenotypic spectrum of pyridoxine-dependent epilepsy (ALDH7A1
deficiency)" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2892945). Brain. 133 (Pt 7):
2148–59. doi:10.1093/brain/awq143 (https://doi.org/10.1093%2Fbrain%2Fawq143).
PMC 2892945 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2892945). PMID 20554659
(https://www.ncbi.nlm.nih.gov/pubmed/20554659).
94. Hagen J, te Brinke H, Wanders RJ, Knegt AC, Oussoren E, Hoogeboom AJ, Ruijter GJ,
Becker D, Schwab KO, Franke I, Duran M, Waterham HR, Sass JO, Houten SM
(September 2015). "Genetic basis of alpha-aminoadipic and alpha-ketoadipic aciduria".
Journal of Inherited Metabolic Disease. 38 (5): 873–9. doi:10.1007/s10545-015-9841-9 (http
s://doi.org/10.1007%2Fs10545-015-9841-9). PMID 25860818 (https://www.ncbi.nlm.nih.gov/
pubmed/25860818).
95. Hedlund GL, Longo N, Pasquali M (May 2006). "Glutaric acidemia type 1" (https://www.ncbi.
nlm.nih.gov/pmc/articles/PMC2556991). American Journal of Medical Genetics Part C:
Seminars in Medical Genetics. 142C (2): 86–94. doi:10.1002/ajmg.c.30088 (https://doi.org/1
0.1002%2Fajmg.c.30088). PMC 2556991 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2
556991). PMID 16602100 (https://www.ncbi.nlm.nih.gov/pubmed/16602100).
96. "Hyperlysinuria | Define Hyperlysinuria at Dictionary.com" (http://dictionary.reference.com/br
owse/hyperlysinuria).
97. Walter, John; John Fernandes; Jean-Marie Saudubray; Georges van den Berghe (2006).
Inborn Metabolic Diseases: Diagnosis and Treatment. Berlin: Springer. p. 296. ISBN 978-3-
540-28783-4.
98. "Norwegian granted for improving lysine production process" (https://web.archive.org/web/2
0120311103403/http://www.allaboutfeed.net/news/norwegian-granted-for-improving-lysine-p
roduction-process-id4052.html). All About Feed. 26 January 2010. Archived from the
original (http://www.allaboutfeed.net/news/norwegian-granted-for-improving-lysine-productio
n-process-id4052.html) on 11 March 2012.
99. Toride Y (2004). "Lysine and other amino acids for feed: production and contribution to
protein utilization in animal feeding" (http://www.fao.org/docrep/007/y5019e/y5019e0a.htm).
Protein sources for the animal feed industry; FAO Expert Consultation and Workshop on
Protein Sources for the Animal Feed Industry; Bangkok, 29 April - 3 May 2002. Rome: Food
and Agriculture Organization of the United Nations. ISBN 978-92-5-105012-5.
100. Abelson PH (March 1999). "A potential phosphate crisis". Science. 283 (5410): 2015.
doi:10.1126/science.283.5410.2015 (https://doi.org/10.1126%2Fscience.283.5410.2015).
PMID 10206902 (https://www.ncbi.nlm.nih.gov/pubmed/10206902).
101. Coyne JA (10 October 1999). "The Truth Is Way Out There" (https://query.nytimes.com/gst/f
ullpage.html?res=9E0DE6D7153EF933A25753C1A96F958260). The New York Times.
Retrieved 6 April 2008.
102. Wu G (May 2009). "Amino acids: metabolism, functions, and nutrition". Amino Acids. 37 (1):
1–17. doi:10.1007/s00726-009-0269-0 (https://doi.org/10.1007%2Fs00726-009-0269-0).
PMID 19301095 (https://www.ncbi.nlm.nih.gov/pubmed/19301095).
103. Connor JM (2008). Global Price Fixing (2nd ed.). Heidelberg: Springer-Verlag. ISBN 978-3-
540-78669-6.
104. Eichenwald K (2000). The Informant: a true story (https://archive.org/details/informanttruest
00eich). New York: Broadway Books. ISBN 978-0-7679-0326-4.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Lysine&oldid=932506631"

This page was last edited on 26 December 2019, at 12:00 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

S-ar putea să vă placă și