Dihydrofolate reductase

Last updated

DHFR
Dihydrofolate reductase 1DRF.png
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases DHFR , DHFRP1, DYR, dihydrofolate reductase
External IDs OMIM: 126060 MGI: 94890 HomoloGene: 56470 GeneCards: DHFR
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_000791
NM_001290354
NM_001290357

NM_010049

RefSeq (protein)

NP_000782
NP_001277283
NP_001277286

NP_034179

Location (UCSC) Chr 5: 80.63 – 80.65 Mb Chr 13: 92.49 – 92.53 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Dihydrofolate reductase, or DHFR, is an enzyme that reduces dihydrofolic acid to tetrahydrofolic acid, using NADPH as an electron donor, which can be converted to the kinds of tetrahydrofolate cofactors used in 1-carbon transfer chemistry. In humans, the DHFR enzyme is encoded by the DHFR gene. [5] [6] It is found in the q14.1 region of chromosome 5. [7]

There are two structural classes of DHFR, evolutionarily unrelated to each other. The former is usually just called DHFR and is found in bacterial chromosomes and animals. In bacteria, however, antibiotic pressure has caused this class to evolve different patterns of binding diaminoheterocyclic molecules, leading to many "types" named under this class, while mammalian ones remain highly similar. [8] The latter (type II), represented by the plastid-encoded R67, is a tiny enzyme that works by forming a homotetramer. [9]

Function

Dihydrofolate reductase
Identifiers
EC no. 1.5.1.3
CAS no. 9002-03-3
Databases
IntEnz IntEnz view
BRENDA BRENDA entry
ExPASy NiceZyme view
KEGG KEGG entry
MetaCyc metabolic pathway
PRIAM profile
PDB structures RCSB PDB PDBe PDBsum
Gene Ontology AmiGO / QuickGO
Search
PMC articles
PubMed articles
NCBI proteins

Dihydrofolate reductase converts dihydrofolate into tetrahydrofolate, a proton shuttle required for the de novo synthesis of purines, thymidylic acid, and certain amino acids. While the functional dihydrofolate reductase gene has been mapped to chromosome 5, multiple intronless processed pseudogenes or dihydrofolate reductase-like genes have been identified on separate chromosomes. [10]

Found in all organisms, DHFR has a critical role in regulating the amount of tetrahydrofolate in the cell. Tetrahydrofolate and its derivatives are essential for purine and thymidylate synthesis, which are important for cell proliferation and cell growth. [11] DHFR plays a central role in the synthesis of nucleic acid precursors, and it has been shown that mutant cells that completely lack DHFR require glycine, a purine, and thymidine to grow. [12] DHFR has also been demonstrated as an enzyme involved in the salvage of tetrahydrobiopterin from dihydrobiopterin [13]


Structure

Dihydrofolate reductase
PDB 8dfr EBI.jpg
Crystal structure of chicken liver dihydrofolate reductase. PDB entry 8dfr
Identifiers
SymbolDHFR_1
Pfam PF00186
Pfam clan CL0387
InterPro IPR001796
PROSITE PDOC00072
SCOP2 1dhi / SCOPe / SUPFAM
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary

A central eight-stranded beta-pleated sheet makes up the main feature of the polypeptide backbone folding of DHFR. [14] Seven of these strands are parallel and the eighth runs antiparallel. Four alpha helices connect successive beta strands. [15] Residues 9 – 24 are termed "Met20" or "loop 1" and, along with other loops, are part of the major subdomain that surround the active site. [16] The active site is situated in the N-terminal half of the sequence, which includes a conserved Pro-Trp dipeptide; the tryptophan has been shown to be involved in the binding of substrate by the enzyme. [15]

Mechanism

General mechanism

The reduction of dihydrofolate to tetrahydrofolate catalyzed by DHFR. DHFR Reaction Scheme.png
The reduction of dihydrofolate to tetrahydrofolate catalyzed by DHFR.

DHFR catalyzes the transfer of a hydride from NADPH to dihydrofolate with an accompanying protonation to produce tetrahydrofolate. [11] In the end, dihydrofolate is reduced to tetrahydrofolate and NADPH is oxidized to NADP+. The high flexibility of Met20 and other loops near the active site play a role in promoting the release of the product, tetrahydrofolate. In particular the Met20 loop helps stabilize the nicotinamide ring of the NADPH to promote the transfer of the hydride from NADPH to dihydrofolate. [16]

The mechanism of this enzyme is stepwise and steady-state random. Specifically, the catalytic reaction begins with the NADPH and the substrate attaching to the binding site of the enzyme, followed by the protonation and the hydride transfer from the cofactor NADPH to the substrate. However, two latter steps do not take place simultaneously in a same transition state. [17] [18] In a study using computational and experimental approaches, Liu et al conclude that the protonation step precedes the hydride transfer. [19]

DHFR (Met20 loop highlighted) + NADPH + folate DHFR + NADPH + folate (Met20 loop).png
DHFR (Met20 loop highlighted) + NADPH + folate

DHFR's enzymatic mechanism is shown to be pH dependent, particularly the hydride transfer step, since pH changes are shown to have remarkable influence on the electrostatics of the active site and the ionization state of its residues. [19] The acidity of the targeted nitrogen on the substrate is important in the binding of the substrate to the enzyme's binding site which is proved to be hydrophobic even though it has direct contact to water. [17] [20] Asp27 is the only charged hydrophilic residue in the binding site, and neutralization of the charge on Asp27 may alter the pKa of the enzyme. Asp27 plays a critical role in the catalytic mechanism by helping with protonation of the substrate and restraining the substrate in the conformation favorable for the hydride transfer. [21] [17] [20] The protonation step is shown to be associated with enol tautomerization even though this conversion is not considered favorable for the proton donation. [18] A water molecule is proved to be involved in the protonation step. [22] [23] [24] Entry of the water molecule to the active site of the enzyme is facilitated by the Met20 loop. [25]

Conformational changes of DHFR

The closed structure is depicted in red and the occluded structure is depicted in green in the catalytic scheme. In the structure, DHF and THF are colored red, NADPH is colored yellow, and Met20 residue is colored blue Conformational changes during the DHFR catalytic cycle.png
The closed structure is depicted in red and the occluded structure is depicted in green in the catalytic scheme. In the structure, DHF and THF are colored red, NADPH is colored yellow, and Met20 residue is colored blue

The catalytic cycle of the reaction catalyzed by DHFR incorporates five important intermediate: holoenzyme (E:NADPH), Michaelis complex (E:NADPH:DHF), ternary product complex (E:NADP+:THF), tetrahydrofolate binary complex (E:THF), and THF‚NADPH complex (E:NADPH:THF). The product (THF) dissociation step from E:NADPH:THF to E:NADPH is the rate determining step during steady-state turnover. [21]

Conformational changes are critical in DHFR's catalytic mechanism. [26] The Met20 loop of DHFR is able to open, close or occlude the active site. [23] [17] Correspondingly, three different conformations classified as the opened, closed and occluded states are assigned to Met20. In addition, an extra distorted conformation of Met20 was defined due to its indistinct characterization results. [23] The Met20 loop is observed in its occluded conformation in the three product ligating intermediates, where the nicotinamide ring is occluded from the active site. This conformational feature accounts for the fact that the substitution of NADP+ by NADPH is prior to product dissociation. Thus, the next round of reaction can occur upon the binding of substrate. [21]

R67 DHFR

R67 dihydrofolate reductase
PDB 2rk1 EBI.png
R67 in complex with DHF and NADP+, monomer. PDB entry 2rk1 .
Identifiers
SymbolDHFR_2
Pfam PF06442
InterPro IPR009159
SCOP2 1vif / SCOPe / SUPFAM
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary

Due to its unique structure and catalytic features, R67 DHFR is widely studied. R67 DHFR is a type II R-plasmid-encoded DHFR without geneticay or structural relation to the E. coli chromosomal DHFR. It is a homotetramer that possesses the 222 symmetry with a single active site pore that is exposed to solvent. [27] This symmetry of active site results in the different binding mode of the enzyme: It can bind with two dihydrofolate (DHF) molecules with positive cooperativity or two NADPH molecules with negative cooperativity, or one substrate plus one, but only the latter one has the catalytical activity. [28] Compare with E. coli chromosomal DHFR, it has higher Km in binding dihydrofolate (DHF) and NADPH. The much lower catalytical kinetics show that hydride transfer is the rate determine step rather than product (THF) release. [29]

In the R67 DHFR structure, the homotetramer forms an active site pore. In the catalytical process, DHF and NADPH enters into the pore from opposite position. The π-π stacking interaction between NADPH's nicotinamide ring and DHF's pteridine ring tightly connect two reactants in the active site. However, the flexibility of p-aminobenzoylglutamate tail of DHF was observed upon binding which can promote the formation of the transition state. [30]

Clinical significance

DHFR mutations cause dihydrofolate reductase deficiency, a rare autosomal recessive inborn error of folate metabolism that results in megaloblastic anemia, pancytopenia and severe cerebral folate deficiency. These issues can be overcome by supplementation with a reduced form of folate, usually folinic acid. [10] [31] [32]

Therapeutic applications

DHFR is an attractive pharmaceutical target for inhibition due to its pivotal role in DNA precursor (thymine) synthesis. Trimethoprim, an antibiotic, inhibits bacterial DHFR while methotrexate, a chemotherapy agent, inhibits mammalian DHFR. However, resistance has developed against some drugs, as a result of mutational changes in DHFR itself. [33]

Cancer

DHFR is responsible for the levels of tetrahydrofolate in a cell, and the inhibition of DHFR can limit the growth and proliferation of cells that are characteristic of cancer and bacterial infections. Methotrexate, a competitive inhibitor of DHFR, is one such anticancer drug that inhibits DHFR. [34]

Folate is necessary for growth, [35] and the pathway of the metabolism of folate is a target in developing treatments for cancer. DHFR is one such target. A regimen of fluorouracil, doxorubicin, and methotrexate was shown to prolong survival in patients with advanced gastric cancer. [36] Further studies into inhibitors of DHFR can lead to more ways to treat cancer.

Infection

Bacteria also need DHFR to grow and multiply and hence inhibitors selective for bacterial DHFR have found application as antibacterial agents. [37] Trimethoprim has shown to have activity against a variety of Gram-positive bacterial pathogens. [37] However, resistance to trimethoprim and other drugs aimed at DHFR can arise due to a variety of mechanisms, limiting the success of their therapeutical uses. [38] [39] [40] Resistance can arise from DHFR gene amplification, mutations in DHFR, [41] [42] decrease in the uptake of the drugs, among others. Regardless, trimethoprim and sulfamethoxazole in combination has been used as an antibacterial agent for decades. [37]

Pyrimethamine is a widely used antiprotozoal agent. [43]

Other classes of compounds that target DHFR in general, and bacterial DHFRs in particular, belong to the classes such as diaminopteridines, diaminotriazines, diaminopyrroloquinazolines, stilbenes, chalcones, deoxybenzoins, diaminoquinazolines, diaminopyrroloquinazolines, to name but a few.

Potential anthrax treatment

Structural alignment of chromosomal (Type I) dihydrofolate reductase from Bacillus anthracis (BaDHFR), Staphylococcus aureus (SaDHFR), Escherichia coli (EcDHFR), and Streptococcus pneumoniae (SpDHFR). Structural alignment of ba sa ec sp dhfr.png
Structural alignment of chromosomal (Type I) dihydrofolate reductase from Bacillus anthracis (BaDHFR), Staphylococcus aureus (SaDHFR), Escherichia coli (EcDHFR), and Streptococcus pneumoniae (SpDHFR).

Dihydrofolate reductase from Bacillus anthracis (BaDHFR) is a validated drug target in the treatment of the infectious disease, anthrax. BaDHFR is less sensitive to trimethoprim analogs than is dihydrofolate reductase from other species such as Escherichia coli , Staphylococcus aureus , and Streptococcus pneumoniae . A structural alignment of dihydrofolate reductase from all four species shows that only BaDHFR has the combination phenylalanine and tyrosine in positions 96 and 102, respectively.

BaDHFR's resistance to trimethoprim analogs is due to these two residues (F96 and Y102), which also confer improved kinetics and catalytic efficiency. [44] Current research uses active site mutants in BaDHFR to guide lead optimization for new antifolate inhibitors. [44]

As a research tool

DHFR has been used as a tool to detect protein–protein interactions in a protein-fragment complementation assay (PCA), using a split-protein approach. [45]

DHFR-lacking CHO cells are the most commonly used cell line for the production of recombinant proteins. These cells are transfected with a plasmid carrying the dhfr gene and the gene for the recombinant protein in a single expression system, and then subjected to selective conditions in thymidine-lacking medium. Only the cells with the exogenous DHFR gene along with the gene of interest survive. Supplementation of this medium with methotrexate, a competitive inhibitor of DHFR, can further select for those cells expressing the highest levels of DHFR, and thus, select for the top recombinant protein producers. [46]

Interactions

Dihydrofolate reductase has been shown to interact with GroEL [47] and Mdm2. [48]

Interactive pathway map

Click on genes, proteins and metabolites below to link to respective articles. [§ 1]

[[File:
FluoropyrimidineActivity WP1601.png go to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to WikiPathwaysgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to article
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
[[
]]
FluoropyrimidineActivity WP1601.png go to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to pathway articlego to pathway articlego to articlego to articlego to articlego to articlego to articlego to WikiPathwaysgo to articlego to articlego to articlego to articlego to articlego to articlego to articlego to articlego to article
|alt=Fluorouracil (5-FU) Activity edit]]
Fluorouracil (5-FU) Activity edit
  1. The interactive pathway map can be edited at WikiPathways: "FluoropyrimidineActivity_WP1601".

Related Research Articles

<span class="mw-page-title-main">Trimethoprim</span> Antibiotic

Trimethoprim (TMP) is an antibiotic used mainly in the treatment of bladder infections. Other uses include for middle ear infections and travelers' diarrhea. With sulfamethoxazole or dapsone it may be used for Pneumocystis pneumonia in people with HIV/AIDS. It is taken orally.

<span class="mw-page-title-main">Methotrexate</span> Chemotherapy and immunosuppressant medication

Methotrexate (MTX), formerly known as amethopterin, is a chemotherapy agent and immune-system suppressant. It is used to treat cancer, autoimmune diseases, and ectopic pregnancies. Types of cancers it is used for include breast cancer, leukemia, lung cancer, lymphoma, gestational trophoblastic disease, and osteosarcoma. Types of autoimmune diseases it is used for include psoriasis, rheumatoid arthritis, and Crohn's disease. It can be given by mouth or by injection.

<span class="mw-page-title-main">Binding site</span> Molecule-specific coordinate bonding area in biological systems

In biochemistry and molecular biology, a binding site is a region on a macromolecule such as a protein that binds to another molecule with specificity. The binding partner of the macromolecule is often referred to as a ligand. Ligands may include other proteins, enzyme substrates, second messengers, hormones, or allosteric modulators. The binding event is often, but not always, accompanied by a conformational change that alters the protein's function. Binding to protein binding sites is most often reversible, but can also be covalent reversible or irreversible.

<span class="mw-page-title-main">Ribonucleotide reductase</span> Class of enzymes

Ribonucleotide reductase (RNR), also known as ribonucleoside diphosphate reductase (rNDP), is an enzyme that catalyzes the formation of deoxyribonucleotides from ribonucleotides. It catalyzes this formation by removing the 2'-hydroxyl group of the ribose ring of nucleoside diphosphates. This reduction produces deoxyribonucleotides. Deoxyribonucleotides in turn are used in the synthesis of DNA. The reaction catalyzed by RNR is strictly conserved in all living organisms. Furthermore, RNR plays a critical role in regulating the total rate of DNA synthesis so that DNA to cell mass is maintained at a constant ratio during cell division and DNA repair. A somewhat unusual feature of the RNR enzyme is that it catalyzes a reaction that proceeds via a free radical mechanism of action. The substrates for RNR are ADP, GDP, CDP and UDP. dTDP is synthesized by another enzyme from dTMP.

<span class="mw-page-title-main">Folinic acid</span> Derivative of folic acid used in cancer treatment

Folinic acid, also known as leucovorin, is a medication used to decrease the toxic effects of methotrexate and pyrimethamine. It is also used in combination with 5-fluorouracil to treat colorectal cancer and pancreatic cancer, may be used to treat folate deficiency that results in anemia, and methanol poisoning. It is taken by mouth, injection into a muscle, or injection into a vein.

Thioredoxin reductases are enzymes that reduce thioredoxin (Trx). Two classes of thioredoxin reductase have been identified: one class in bacteria and some eukaryotes and one in animals. In bacteria TrxR also catalyzes the reduction of glutaredoxin like proteins known as NrdH. Both classes are flavoproteins which function as homodimers. Each monomer contains a FAD prosthetic group, a NADPH binding domain, and an active site containing a redox-active disulfide bond.

<span class="mw-page-title-main">Methionine synthase</span> Mammalian protein found in Homo sapiens

Methionine synthase (MS, MeSe, MTR) is responsible for the regeneration of methionine from homocysteine. In humans it is encoded by the MTR gene (5-methyltetrahydrofolate-homocysteine methyltransferase). Methionine synthase forms part of the S-adenosylmethionine (SAMe) biosynthesis and regeneration cycle, and is the enzyme responsible for linking the cycle to one-carbon metabolism via the folate cycle. There are two primary forms of this enzyme, the Vitamin B12 (cobalamin)-dependent (MetH) and independent (MetE) forms, although minimal core methionine synthases that do not fit cleanly into either category have also been described in some anaerobic bacteria. The two dominant forms of the enzymes appear to be evolutionary independent and rely on considerably different chemical mechanisms. Mammals and other higher eukaryotes express only the cobalamin-dependent form. In contrast, the distribution of the two forms in Archaeplastida (plants and algae) is more complex. Plants exclusively possess the cobalamin-independent form, while algae have either one of the two, depending on species. Many different microorganisms express both the cobalamin-dependent and cobalamin-independent forms.

<span class="mw-page-title-main">Filamentation</span>

Filamentation is the anomalous growth of certain bacteria, such as Escherichia coli, in which cells continue to elongate but do not divide. The cells that result from elongation without division have multiple chromosomal copies.

<span class="mw-page-title-main">Enzyme inhibitor</span> Molecule that blocks enzyme activity

An enzyme inhibitor is a molecule that binds to an enzyme and blocks its activity. Enzymes are proteins that speed up chemical reactions necessary for life, in which substrate molecules are converted into products. An enzyme facilitates a specific chemical reaction by binding the substrate to its active site, a specialized area on the enzyme that accelerates the most difficult step of the reaction.

<span class="mw-page-title-main">Aldose reductase</span> Enzyme

In enzymology, aldose reductase is a cytosolic NADPH-dependent oxidoreductase that catalyzes the reduction of a variety of aldehydes and carbonyls, including monosaccharides. It is primarily known for catalyzing the reduction of glucose to sorbitol, the first step in polyol pathway of glucose metabolism.

<span class="mw-page-title-main">Thymidylate synthase</span> Enzyme

Thymidylate synthase (TS) is an enzyme that catalyzes the conversion of deoxyuridine monophosphate (dUMP) to deoxythymidine monophosphate (dTMP). Thymidine is one of the nucleotides in DNA. With inhibition of TS, an imbalance of deoxynucleotides and increased levels of dUMP arise. Both cause DNA damage.

<span class="mw-page-title-main">Tetrahydrofolic acid</span> Chemical compound

Tetrahydrofolic acid (THFA), or tetrahydrofolate, is a folic acid derivative.

<span class="mw-page-title-main">Serine hydroxymethyltransferase</span>

Serine hydroxymethyltransferase (SHMT) is a pyridoxal phosphate (PLP) (Vitamin B6) dependent enzyme (EC 2.1.2.1) which plays an important role in cellular one-carbon pathways by catalyzing the reversible, simultaneous conversions of L-serine to glycine and tetrahydrofolate (THF) to 5,10-Methylenetetrahydrofolate (5,10-CH2-THF). This reaction provides the largest part of the one-carbon units available to the cell.

<span class="mw-page-title-main">2,4 Dienoyl-CoA reductase</span> Class of enzymes

2,4 Dienoyl-CoA reductase also known as DECR1 is an enzyme which in humans is encoded by the DECR1 gene which resides on chromosome 8. This enzyme catalyzes the following reactions

Flavin reductase a class of enzymes. There are a variety of flavin reductases, which bind free flavins and through hydrogen bonding, catalyze the reduction of these molecules to a reduced flavin. Riboflavin, or vitamin B, and flavin mononucleotide are two of the most well known flavins in the body and are used in a variety of processes which include metabolism of fat and ketones and the reduction of methemoglobin in erythrocytes. Flavin reductases are similar and often confused for ferric reductases because of their similar catalytic mechanism and structures.

<span class="mw-page-title-main">Antifolate</span> Class of antimetabolite medications

Antifolates are a class of antimetabolite medications that antagonise (that is, block) the actions of folic acid (vitamin B9). Folic acid's primary function in the body is as a cofactor to various methyltransferases involved in serine, methionine, thymidine and purine biosynthesis. Consequently, antifolates inhibit cell division, DNA/RNA synthesis and repair and protein synthesis. Some such as proguanil, pyrimethamine and trimethoprim selectively inhibit folate's actions in microbial organisms such as bacteria, protozoa and fungi. The majority of antifolates work by inhibiting dihydrofolate reductase (DHFR).

<span class="mw-page-title-main">Dihydrofolate reductase inhibitor</span> Cellular enzyme inhibitor

A dihydrofolate reductase inhibitor is a molecule that inhibits the function of dihydrofolate reductase, and is a type of antifolate.

<span class="mw-page-title-main">Phosphoribosylglycinamide formyltransferase</span>

Phosphoribosylglycinamide formyltransferase (EC 2.1.2.2, 2-amino-N-ribosylacetamide 5'-phosphate transformylase, GAR formyltransferase, GAR transformylase, glycinamide ribonucleotide transformylase, GAR TFase, 5,10-methenyltetrahydrofolate:2-amino-N-ribosylacetamide ribonucleotide transformylase) is an enzyme with systematic name 10-formyltetrahydrofolate:5'-phosphoribosylglycinamide N-formyltransferase. This enzyme catalyses the following chemical reaction

<span class="mw-page-title-main">Stephen J. Benkovic</span> American chemist

Stephen James Benkovic is an American chemist known for his contributions to the field of enzymology. He holds the Evan Pugh University Professorship and Eberly Chair in Chemistry at The Pennsylvania State University. He has developed boron compounds that are active pharmacophores against a variety of diseases. Benkovic has concentrated on the assembly and kinetic attributes of the enzymatic machinery that performs DNA replication, DNA repair, and purine biosynthesis.

<span class="mw-page-title-main">Mercury(II) reductase</span>

Mercury(II) reductase (EC 1.16.1.1), commonly known as MerA, is an oxidoreductase enzyme and flavoprotein that catalyzes the reduction of Hg2+ to Hg0. Mercury(II) reductase is found in the cytoplasm of many eubacteria in both aerobic and anaerobic environments and serves to convert toxic mercury ions into relatively inert elemental mercury.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000228716 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000021707 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. Chen MJ, Shimada T, Moulton AD, Harrison M, Nienhuis AW (December 1982). "Intronless human dihydrofolate reductase genes are derived from processed RNA molecules". Proceedings of the National Academy of Sciences of the United States of America. 79 (23): 7435–9. Bibcode:1982PNAS...79.7435C. doi: 10.1073/pnas.79.23.7435 . PMC   347354 . PMID   6961421.
  6. Chen MJ, Shimada T, Moulton AD, Cline A, Humphries RK, Maizel J, Nienhuis AW (March 1984). "The functional human dihydrofolate reductase gene". The Journal of Biological Chemistry. 259 (6): 3933–43. doi: 10.1016/S0021-9258(17)43186-3 . PMID   6323448.
  7. "DHFR dihydrofolate reductase [Homo sapiens (human)]". Gene - NCBI. Retrieved 21 February 2023.
  8. Smith SL, Patrick P, Stone D, Phillips AW, Burchall JJ (November 1979). "Porcine liver dihydrofolate reductase. Purification, properties, and amino acid sequence". The Journal of Biological Chemistry. 254 (22): 11475–84. doi: 10.1016/S0021-9258(19)86510-9 . PMID   500653.
  9. Krahn JM, Jackson MR, DeRose EF, Howell EE, London RE (25 December 2007). "Crystal structure of a type II dihydrofolate reductase catalytic ternary complex". Biochemistry. 46 (51): 14878–88. doi:10.1021/bi701532r. PMC   3743094 . PMID   18052202.
  10. 1 2 "Entrez Gene: DHFR dihydrofolate reductase".
  11. 1 2 Schnell JR, Dyson HJ, Wright PE (2004). "Structure, dynamics, and catalytic function of dihydrofolate reductase". Annual Review of Biophysics and Biomolecular Structure. 33 (1): 119–40. doi:10.1146/annurev.biophys.33.110502.133613. PMID   15139807. S2CID   28611812.
  12. Urlaub G, Chasin LA (July 1980). "Isolation of Chinese hamster cell mutants deficient in dihydrofolate reductase activity". Proceedings of the National Academy of Sciences of the United States of America. 77 (7): 4216–20. Bibcode:1980PNAS...77.4216U. doi: 10.1073/pnas.77.7.4216 . PMC   349802 . PMID   6933469.
  13. Crabtree MJ, Tatham AL, Hale AB, Alp NJ, Channon KM (October 2009). "Critical role for tetrahydrobiopterin recycling by dihydrofolate reductase in regulation of endothelial nitric-oxide synthase coupling: relative importance of the de novo biopterin synthesis versus salvage pathways". The Journal of Biological Chemistry. 284 (41): 28128–36. doi: 10.1074/jbc.M109.041483 . PMC   2788863 . PMID   19666465.
  14. Matthews DA, Alden RA, Bolin JT, Freer ST, Hamlin R, Xuong N, Kraut J, Poe M, Williams M, Hoogsteen K (July 1977). "Dihydrofolate reductase: x-ray structure of the binary complex with methotrexate". Science. 197 (4302): 452–5. Bibcode:1977Sci...197..452M. doi:10.1126/science.17920. PMID   17920.
  15. 1 2 Filman DJ, Bolin JT, Matthews DA, Kraut J (November 1982). "Crystal structures of Escherichia coli and Lactobacillus casei dihydrofolate reductase refined at 1.7 Å resolution. II. Environment of bound NADPH and implications for catalysis". The Journal of Biological Chemistry. 257 (22): 13663–72. doi: 10.1016/S0021-9258(18)33498-7 . PMID   6815179.
  16. 1 2 Osborne MJ, Schnell J, Benkovic SJ, Dyson HJ, Wright PE (August 2001). "Backbone dynamics in dihydrofolate reductase complexes: role of loop flexibility in the catalytic mechanism". Biochemistry. 40 (33): 9846–59. doi:10.1021/bi010621k. PMID   11502178.
  17. 1 2 3 4 Rod TH, Brooks CL (July 2003). "How dihydrofolate reductase facilitates protonation of dihydrofolate". Journal of the American Chemical Society. 125 (29): 8718–9. doi:10.1021/ja035272r. PMID   12862454.
  18. 1 2 Wan Q, Bennett BC, Wilson MA, Kovalevsky A, Langan P, Howell EE, Dealwis C (December 2014). "Toward resolving the catalytic mechanism of dihydrofolate reductase using neutron and ultrahigh-resolution X-ray crystallography". Proceedings of the National Academy of Sciences of the United States of America. 111 (51): 18225–30. Bibcode:2014PNAS..11118225W. doi: 10.1073/pnas.1415856111 . PMC   4280638 . PMID   25453083.
  19. 1 2 Liu CT, Francis K, Layfield JP, Huang X, Hammes-Schiffer S, Kohen A, Benkovic SJ (December 2014). "Escherichia coli dihydrofolate reductase catalyzed proton and hydride transfers: temporal order and the roles of Asp27 and Tyr100". Proceedings of the National Academy of Sciences of the United States of America. 111 (51): 18231–6. Bibcode:2014PNAS..11118231L. doi: 10.1073/pnas.1415940111 . PMC   4280594 . PMID   25453098.
  20. 1 2 Czekster CM, Vandemeulebroucke A, Blanchard JS (January 2011). "Kinetic and chemical mechanism of the dihydrofolate reductase from Mycobacterium tuberculosis". Biochemistry. 50 (3): 367–75. doi:10.1021/bi1016843. PMC   3074011 . PMID   21138249.
  21. 1 2 3 Fierke CA, Johnson KA, Benkovic SJ (June 1987). "Construction and evaluation of the kinetic scheme associated with dihydrofolate reductase from Escherichia coli". Biochemistry. 26 (13): 4085–92. doi:10.1021/bi00387a052. PMID   3307916.
  22. Reyes VM, Sawaya MR, Brown KA, Kraut J (February 1995). "Isomorphous crystal structures of Escherichia coli dihydrofolate reductase complexed with folate, 5-deazafolate, and 5,10-dideazatetrahydrofolate: mechanistic implications". Biochemistry. 34 (8): 2710–23. doi:10.1021/bi00008a039. PMID   7873554.
  23. 1 2 3 Sawaya MR, Kraut J (January 1997). "Loop and subdomain movements in the mechanism of Escherichia coli dihydrofolate reductase: crystallographic evidence". Biochemistry. 36 (3): 586–603. doi:10.1021/bi962337c. PMID   9012674.
  24. Chen YQ, Kraut J, Blakley RL, Callender R (June 1994). "Determination by Raman spectroscopy of the pKa of N5 of dihydrofolate bound to dihydrofolate reductase: mechanistic implications". Biochemistry. 33 (23): 7021–6. doi:10.1021/bi00189a001. PMID   8003467.
  25. Shrimpton P, Allemann RK (June 2002). "Role of water in the catalytic cycle of E. coli dihydrofolate reductase". Protein Science. 11 (6): 1442–51. doi:10.1110/ps.5060102. PMC   2373639 . PMID   12021443.
  26. Antikainen NM, Smiley RD, Benkovic SJ, Hammes GG (December 2005). "Conformation coupled enzyme catalysis: single-molecule and transient kinetics investigation of dihydrofolate reductase". Biochemistry. 44 (51): 16835–43. doi:10.1021/bi051378i. PMID   16363797.
  27. Narayana N, Matthews DA, Howell EE, Nguyen-huu X (November 1995). "A plasmid-encoded dihydrofolate reductase from trimethoprim-resistant bacteria has a novel D2-symmetric active site". Nature Structural Biology. 2 (11): 1018–25. doi:10.1038/nsb1195-1018. PMID   7583655. S2CID   11914241.
  28. Bradrick TD, Beechem JM, Howell EE (September 1996). "Unusual binding stoichiometries and cooperativity are observed during binary and ternary complex formation in the single active pore of R67 dihydrofolate reductase, a D2 symmetric protein". Biochemistry. 35 (35): 11414–24. doi:10.1021/bi960205d. PMID   8784197.
  29. Park H, Zhuang P, Nichols R, Howell EE (January 1997). "Mechanistic studies of R67 dihydrofolate reductase. Effects of pH and an H62C mutation". The Journal of Biological Chemistry. 272 (4): 2252–8. doi: 10.1074/jbc.272.4.2252 . PMID   8999931.
  30. Kamath G, Howell EE, Agarwal PK (October 2010). "The tail wagging the dog: insights into catalysis in R67 dihydrofolate reductase". Biochemistry. 49 (42): 9078–88. doi:10.1021/bi1007222. PMID   20795731.
  31. Banka S, Blom HJ, Walter J, Aziz M, Urquhart J, Clouthier CM, Rice GI, de Brouwer AP, Hilton E, Vassallo G, Will A, Smith DE, Smulders YM, Wevers RA, Steinfeld R, Heales S, Crow YJ, Pelletier JN, Jones S, Newman WG (February 2011). "Identification and characterization of an inborn error of metabolism caused by dihydrofolate reductase deficiency". American Journal of Human Genetics. 88 (2): 216–25. doi:10.1016/j.ajhg.2011.01.004. PMC   3035707 . PMID   21310276.
  32. Nyhan WL, Hoffmann GF, Barshop BA (30 December 2011). Atlas of Inherited Metabolic Diseases 3E. CRC Press. pp. 141–. ISBN   978-1-4441-4948-7.
  33. Cowman AF, Lew AM (November 1989). "Antifolate drug selection results in duplication and rearrangement of chromosome 7 in Plasmodium chabaudi". Molecular and Cellular Biology. 9 (11): 5182–8. doi:10.1128/mcb.9.11.5182. PMC   363670 . PMID   2601715.
  34. Li R, Sirawaraporn R, Chitnumsub P, Sirawaraporn W, Wooden J, Athappilly F, Turley S, Hol WG (January 2000). "Three-dimensional structure of M. tuberculosis dihydrofolate reductase reveals opportunities for the design of novel tuberculosis drugs". Journal of Molecular Biology. 295 (2): 307–23. doi:10.1006/jmbi.1999.3328. PMID   10623528. S2CID   24527344.
  35. Bailey SW, Ayling JE (September 2009). "The extremely slow and variable activity of dihydrofolate reductase in human liver and its implications for high folic acid intake". Proceedings of the National Academy of Sciences of the United States of America. 106 (36): 15424–9. doi: 10.1073/pnas.0902072106 . PMC   2730961 . PMID   19706381.
  36. Murad AM, Santiago FF, Petroianu A, Rocha PR, Rodrigues MA, Rausch M (July 1993). "Modified therapy with 5-fluorouracil, doxorubicin, and methotrexate in advanced gastric cancer". Cancer. 72 (1): 37–41. doi: 10.1002/1097-0142(19930701)72:1<37::AID-CNCR2820720109>3.0.CO;2-P . PMID   8508427.
  37. 1 2 3 Hawser S, Lociuro S, Islam K (March 2006). "Dihydrofolate reductase inhibitors as antibacterial agents". Biochemical Pharmacology. 71 (7): 941–8. doi:10.1016/j.bcp.2005.10.052. PMID   16359642.
  38. Narayana N, Matthews DA, Howell EE, Nguyen-huu X (November 1995). "A plasmid-encoded dihydrofolate reductase from trimethoprim-resistant bacteria has a novel D2-symmetric active site". Nature Structural Biology. 2 (11): 1018–25. doi:10.1038/nsb1195-1018. PMID   7583655. S2CID   11914241.
  39. Huennekens FM (June 1996). "In search of dihydrofolate reductase". Protein Science. 5 (6): 1201–8. doi:10.1002/pro.5560050626. PMC   2143423 . PMID   8762155.
  40. Banerjee D, Mayer-Kuckuk P, Capiaux G, Budak-Alpdogan T, Gorlick R, Bertino JR (July 2002). "Novel aspects of resistance to drugs targeted to dihydrofolate reductase and thymidylate synthase". Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 1587 (2–3): 164–73. doi: 10.1016/S0925-4439(02)00079-0 . PMID   12084458.
  41. Toprak E, Veres A, Michel JB, Chait R, Hartl DL, Kishony R (December 2011). "Evolutionary paths to antibiotic resistance under dynamically sustained drug selection". Nature Genetics. 44 (1): 101–5. doi:10.1038/ng.1034. PMC   3534735 . PMID   22179135.
  42. Rodrigues JV, Bershtein S, Li A, Lozovsky ER, Hartl DL, Shakhnovich EI (March 2016). "Biophysical principles predict fitness landscapes of drug resistance". Proceedings of the National Academy of Sciences of the United States of America. 113 (11): E1470-8. Bibcode:2016PNAS..113E1470R. doi: 10.1073/pnas.1601441113 . PMC   4801265 . PMID   26929328.
  43. Benkovic SJ, Fierke CA, Naylor AM (March 1988). "Insights into enzyme function from studies on mutants of dihydrofolate reductase". Science. 239 (4844): 1105–10. Bibcode:1988Sci...239.1105B. doi:10.1126/science.3125607. PMID   3125607.
  44. 1 2 Beierlein JM, Karri NG, Anderson AC (October 2010). "Targeted mutations of Bacillus anthracis dihydrofolate reductase condense complex structure−activity relationships". Journal of Medicinal Chemistry. 53 (20): 7327–36. doi:10.1021/jm100727t. PMC   3618964 . PMID   20882962.
  45. Tarassov K, Messier V, Landry CR, Radinovic S, Serna Molina MM, Shames I, Malitskaya Y, Vogel J, Bussey H, Michnick SW (June 2008). "An in vivo map of the yeast protein interactome" (PDF). Science. 320 (5882): 1465–70. Bibcode:2008Sci...320.1465T. doi:10.1126/science.1153878. PMID   18467557. S2CID   1732896.
  46. Ng SK (2012). "Generation of High-Expressing Cells by Methotrexate Amplification of Destabilized Dihydrofolate Reductase Selection Marker". Protein Expression in Mammalian Cells. Methods in Molecular Biology. Vol. 801. pp. 161–172. doi:10.1007/978-1-61779-352-3_11. ISBN   978-1-61779-351-6. PMID   21987253.
  47. Mayhew M, da Silva AC, Martin J, Erdjument-Bromage H, Tempst P, Hartl FU (February 1996). "Protein folding in the central cavity of the GroEL-GroES chaperonin complex". Nature. 379 (6564): 420–6. Bibcode:1996Natur.379..420M. doi:10.1038/379420a0. PMID   8559246. S2CID   4310511.
  48. Maguire M, Nield PC, Devling T, Jenkins RE, Park BK, Polański R, Vlatković N, Boyd MT (May 2008). "MDM2 regulates dihydrofolate reductase activity through monoubiquitination". Cancer Research. 68 (9): 3232–42. doi:10.1158/0008-5472.CAN-07-5271. PMC   3536468 . PMID   18451149.

Further reading

This article incorporates text from the public domain Pfam and InterPro: IPR001796
This article incorporates text from the public domain Pfam and InterPro: IPR009159