Search results for PRDX1

Showing 17 results out of 20

×

Species

Types

Compartments

Reaction types

Search properties

Species

Types

Compartments

Reaction types

Search properties

Protein (4 results from a total of 4)

Identifier: R-HSA-3341270
Species: Homo sapiens
Compartment: cytosol
Primary external reference: UniProt: PRDX1: Q06830
Identifier: R-HSA-8868564
Species: Homo sapiens
Compartment: cytosol
Primary external reference: UniProt: PRDX1: Q06830
Identifier: R-HSA-5629198
Species: Homo sapiens
Compartment: cytosol
Primary external reference: UniProt: PRDX1: Q06830
Identifier: R-HSA-3323008
Species: Homo sapiens
Compartment: mitochondrial matrix
Primary external reference: UniProt: PRDX5: P30044

DNA Sequence (1 results from a total of 1)

Identifier: R-HSA-8958034
Species: Homo sapiens
Compartment: nucleoplasm
Primary external reference: ENSEMBL: ENSEMBL:ENSG00000117450

Reaction (5 results from a total of 8)

Identifier: R-HSA-5631885
Species: Homo sapiens
Compartment: cytosol
The activity of eukaryotic PRDX1 gradually decreases with time, which is due to the overoxidation of the catalytic cysteine C52. Normally, oxidized cysteine C52-SOH is generated as a catalytic intermediate, which is subsequently reduced by thioredoxin. Occasionally, further oxidation happens, generating C52-SOOH , where the catalytic cysteine is converted to cysteine-sulfinic acid. This over-oxidation cannot be reversed by thioredoxin (Yang et al. 2002, Budanov et al. 2004). Bacterial peroxiredoxin AhpC does not undergo over-oxidation due to structural difference (Wood et al. 2003).
Identifier: R-HSA-8868567
Species: Homo sapiens
Compartment: cytosol
Oxidative stress, manifested through accumulation of reactive oxygen species in the cell, is one of the hallmarks of Alzheimer's disease. Based on mouse model studies, CDK5, aberrantly activated by binding to p25, phosphorylates the peroxide reductase PRDX1 on a conserved threonine residue T90, thus inactivating it and contributing to ROS accumulation (Sun et al. 2008).
Identifier: R-HSA-5631903
Species: Homo sapiens
Compartment: cytosol
Sestrins (SESN1, SESN2 and likely SESN3) bind overoxidized PRDX1, in which the catalytic cysteine C52 has been converted to cysteine-sulfinic acid. Among all peroxiredoxins, PRDX1 is the most abundant member of the PRDX family. The major function is to protect cells against reactive oxygen species (ROS), thus impacting on cell proliferation and survival (Gong et al. 2015). While several reports state that sestrins reduce overoxidized PRDX1 to the catalytically active homodimer (Budanov et al. 2004, Papadia et al. 2008, Essler et al. 2009), there are conflicting reports claiming that sestrins do not possess cysteine sulfinyl reductase activity (Woo et al. 2009).
Identifier: R-HSA-9759129
Species: Homo sapiens
Compartment: nucleoplasm
NFE2L2 binds to its consensus sequence in the promoter of the PRDX1 gene to stimulate expression downstream of hypoxia, antioxidant and electrophilic stressors (Ishii et al, 2000; Kwak et al, 2003; Kim et al, 2007; reviewed in Baird and Yamamoto, 2020). PRDX1 protein contributes to stress response through the reduction of hydrogen peroxide and other substrates (reviewed in Ding et al, 2017; Ledgerwood et al, 2017).
Identifier: R-HSA-9759189
Species: Homo sapiens
Compartment: nucleoplasm, cytosol
PRDX1 encodes a thiol-specific peroxidase that contributes to cytoprotection by reducing hydrogen peroxide and other hydroperoxides to water and alcohols (reviewed in Ding et al, 2017; Ledgerwood et al, 2017). PRDX1 expression is stimulated downstream of the NFE2L2-KEAP1 pathway in response to oxidative and electrophilic stressors (Ishii et al, 2000; Kwak et al, 2003). Consistent with this, NFE2L2 binds to its cognate element in the PRDX1 promoter as assessed by electrophoretic mobility shift (EMSA) assay and chromatin immunoprecipitation (ChIP) in response to hypoxia (Kim et al, 2007; reviewed in Baird and Yamamoto, 2020).

Complex (4 results from a total of 4)

Identifier: R-HSA-3341319
Species: Homo sapiens
Compartment: cytosol
Identifier: R-HSA-9759106
Species: Homo sapiens
Compartment: nucleoplasm
Identifier: R-HSA-5631882
Species: Homo sapiens
Compartment: cytosol
Identifier: R-HSA-5631902
Species: Homo sapiens
Compartment: cytosol

Set (1 results from a total of 1)

Identifier: R-HSA-3341359
Species: Homo sapiens
Compartment: cytosol

Pathway (2 results from a total of 2)

Identifier: R-HSA-9818027
Species: Homo sapiens
Compartment: nucleoplasm
Subpathway representing cytoprotective genes regulated by NFE2L2 (NRF2). NFE2L2 is well-studied for its role in oxidative stress where it gets activated by ROS and then induces a plethora of gene expression regulation the oxidative damage. It induces genes/enzymes that regulate the phase 2 detoxification system (eg. GSTs and Glutathione system), ROS scavenging (SODs,PRDX1 ) and cytoprotection (HO1) by regulating inflammation and tissue damage (Tonelli et al, 2018; Shaw et al, 2020)
Identifier: R-HSA-5628897
Species: Homo sapiens
While the p53 tumor suppressor protein (TP53) is known to inhibit cell growth by inducing apoptosis, senescence and cell cycle arrest, recent studies have found that p53 is also able to influence cell metabolism to prevent tumor development. TP53 regulates transcription of many genes involved in the metabolism of carbohydrates, nucleotides and amino acids, protein synthesis and aerobic respiration.

TP53 stimulates transcription of TIGAR, a D-fructose 2,6-bisphosphatase. TIGAR activity decreases glycolytic rate and lowers ROS (reactive oxygen species) levels in cells (Bensaad et al. 2006). TP53 may also negatively regulate the rate of glycolysis by inhibiting the expression of glucose transporters GLUT1, GLUT3 and GLUT4 (Kondoh et al. 2005, Schwartzenberg-Bar-Yoseph et al. 2004, Kawauchi et al. 2008).

TP53 negatively regulates several key points in PI3K/AKT signaling and downstream mTOR signaling, decreasing the rate of protein synthesis and, hence, cellular growth. TP53 directly stimulates transcription of the tumor suppressor PTEN, which acts to inhibit PI3K-mediated activation of AKT (Stambolic et al. 2001). TP53 stimulates transcription of sestrin genes, SESN1, SESN2, and SESN3 (Velasco-Miguel et al. 1999, Budanov et al. 2002, Brynczka et al. 2007). One of sestrin functions may be to reduce and reactivate overoxidized peroxiredoxin PRDX1, thereby reducing ROS levels (Budanov et al. 2004, Papadia et al. 2008, Essler et al. 2009). Another function of sestrins is to bind the activated AMPK complex and protect it from AKT-mediated inactivation. By enhancing AMPK activity, sestrins negatively regulate mTOR signaling (Budanov and Karin 2008, Cam et al. 2014). The expression of DDIT4 (REDD1), another negative regulator of mTOR signaling, is directly stimulated by TP63 and TP53. DDIT4 prevents AKT-mediated inactivation of TSC1:TSC2 complex, thus inhibiting mTOR cascade (Cam et al. 2014, Ellisen et al. 2002, DeYoung et al. 2008). TP53 may also be involved, directly or indirectly, in regulation of expression of other participants of PI3K/AKT/mTOR signaling, such as PIK3CA (Singh et al. 2002), TSC2 and AMPKB (Feng et al. 2007).

TP53 regulates mitochondrial metabolism through several routes. TP53 stimulates transcription of SCO2 gene, which encodes a mitochondrial cytochrome c oxidase assembly protein (Matoba et al. 2006). TP53 stimulates transcription of RRM2B gene, which encodes a subunit of the ribonucleotide reductase complex, responsible for the conversion of ribonucleotides to deoxyribonucleotides and essential for the maintenance of mitochondrial DNA content in the cell (Tanaka et al. 2000, Bourdon et al. 2007, Kulawiec et al. 2009). TP53 also transactivates mitochondrial transcription factor A (TFAM), a nuclear-encoded gene important for mitochondrial DNA (mtDNA) transcription and maintenance (Park et al. 2009). Finally, TP53 stimulates transcription of the mitochondrial glutaminase GLS2, leading to increased mitochondrial respiration rate and reduced ROS levels (Hu et al. 2010).

The great majority of tumor cells generate energy through aerobic glycolysis, rather than the much more efficient aerobic mitochondrial respiration, and this metabolic change is known as the Warburg effect (Warburg 1956). Since the majority of tumor cells have impaired TP53 function, and TP53 regulates a number of genes involved in glycolysis and mitochondrial respiration, it is likely that TP53 inactivation plays an important role in the metabolic derangement of cancer cells such as the Warburg effect and the concomitant increased tumorigenicity (reviewed by Feng and Levine 2010). On the other hand, some mutations of TP53 in Li-Fraumeni syndrome may result in the retention of its wild-type metabolic activities while losing cell cycle and apoptosis functions (Wang et al. 2013). Consistent with such human data, some mutations of p53, unlike p53 null state, retain the ability to regulate energy metabolism while being inactive in regulating its classic gene targets involved in cell cycle, apoptosis and senescence. Retention of metabolic and antioxidant functions of p53 protects p53 mutant mice from early onset tumorigenesis (Li et al. 2012).

Cite Us!