Cannabis Sativa

SLC46A3
Identifiers
AliasesSLC46A3, FKSG16, SLC46A3 (gene), solute carrier family 46 member 3
External IDsOMIM: 616764 MGI: 1918956 HomoloGene: 41733 GeneCards: SLC46A3
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001135919
NM_181785
NM_001347960

NM_027872
NM_001357002

RefSeq (protein)

NP_001129391
NP_861450
NP_001334889

NP_082148
NP_001343931

Location (UCSC)Chr 13: 28.7 – 28.72 MbChr 5: 147.82 – 147.83 Mb
PubMed search[3][4]
Wikidata
View/Edit HumanView/Edit Mouse

Solute carrier family 46 member 3 (SLC46A3) is a protein that in humans is encoded by the SLC46A3 gene.[5] Also referred to as FKSG16, the protein belongs to the major facilitator superfamily (MFS) and SLC46A family.[6] Most commonly found in the plasma membrane and endoplasmic reticulum (ER), SLC46A3 is a multi-pass membrane protein with 11 α-helical transmembrane domains.[7][8] It is mainly involved in the transport of small molecules across the membrane through the substrate translocation pores featured in the MFS domain.[9][10] The protein is associated with breast and prostate cancer, hepatocellular carcinoma (HCC), papilloma, glioma, obesity, and SARS-CoV.[11][12][13][14][15][16] Based on the differential expression of SLC46A3 in antibody-drug conjugate (ADC)-resistant cells and certain cancer cells, current research is focused on the potential of SLC46A3 as a prognostic biomarker and therapeutic target for cancer.[17] While protein abundance is relatively low in humans, high expression has been detected particularly in the liver, small intestine, and kidney.[18][19]

Gene[edit]

The SLC46A3 gene, also known by its aliases solute carrier family 46 member 3 and FKSG16, is located at 13q12.3 on the reverse strand in humans.[5] The gene spans 18,950 bases from 28,700,064 to 28,719,013 (GRCh38/hg38), flanked by POMP upstream and CYP51A1P2 downstream.[6][20] SLC46A3 contains 6 exons and 5 introns.[5] There are two paralogs for this gene, SLC46A1 and SLC46A2, and orthologs as distant as fungi.[21] So far, more than 4580 single nucleotide polymorphisms (SNPs) for this gene have been identified.[22] SLC46A3 is expressed at relatively low levels, about 0.5x the average gene.[23] Gene expression is peculiarly high in the liver, small intestine, and kidney.[18][19]

Transcript[edit]

Transcript Variants[edit]

SLC46A3 has multiple transcript variants produced by different promoter regions and alternative splicing.[5][24] A total of 4 transcript variants are found in the RefSeq database.[25] Variant 1 is most abundant.[26]

Transcript Variants of SLC46A3
Transcript Variant Accession Number Length (bp) Description
1[26] NM_181785.4 3302 MANE select. Variant 1 encodes isoform a.
2[27] NM_001135919.2 2758 Variant 2 encodes isoform b. It lacks a segment in the 3' coding region and the resulting frameshift causes isoform b to have a longer C-terminus than isoform a.
3[28] NM_001347960.1 3099 Variant 3 also encodes isoform a. Variants 1 and 3 differ in their 5' untranslated regions (UTRs).
X1[29] XM_005266361.2 1845 Variant X1 encodes isoform X1.

*Lengths shown do not include introns.

Protein[edit]

Isoforms[edit]

3 isoforms have been reported for SLC46A3.[5] Isoform a is MANE select and most abundant.[30] All isoforms contain the MFS and MFS_1 domains as well as the 11 transmembrane regions.[8][31][32]

Isoforms of the SLC46A3 Protein
Isoform Accession Number Length (aa) Transcript
a[30][8] NP_861450.1

NP_001334889.1

461 1,3
b[31] NP_001129391.1 463 2
X1[32] XP_005266418.1 463 X1

*Lengths shown are for the precursor proteins.

Properties[edit]

SLC46A3 is an integral membrane protein 461 amino acids (aa) of length with a molecular weight (MW) of 51.5 kDa.[33] The basal isoelectric point (pI) for this protein is 5.56.[34] The protein contains 11 transmembrane domains in addition to domains MFS and MFS_1.[30] MFS and MFS_1 domains largely overlap and contain 42 putative substrate translocation pores that are predicted to bind substrates for transmembrane transport.[10] The substrate translocation pores have access to both sides of the membrane in an alternating fashion through a conformational change. SLC46A3 lacks charged and polar amino acids while containing an excess of nonpolar amino acids, particularly phenylalanine (Phe).[33] The resulting hydrophobicity is mostly concentrated in the transmembrane regions for interactions with the fatty acid chains in the lipid bilayer.[35] The transmembrane domains also have a shortage of proline (Pro), a helix breaker.[33]

SLC46A3 Protein Sequence Analysis. The blue bars enclose the MFS domain and the red brackets the transmembrane regions. i = LVIF.

The protein sequence contains mixed, positive, and negative charge clusters, one of each, which are high in glutamine (Glu).[33] The clusters are located outside the transmembrane regions, and thus are solvent-exposed. Two 0 runs that run through several transmembrane domains in addition to a +/* run in between two transmembrane domains are also present. The protein contains a C-(X)2-C motif (CLLC), which is mostly present in metal-binding proteins and oxidoreductases.[36] A sorting-signal sequence motif, YXXphi, is also found at Tyr246 - Phe249 (YMLF) and Tyr446 - Leu449 (YELL).[37][38] This Y-based sorting signal directs the trafficking within the endosomal and the secretory pathways of integral membrane proteins by interacting with the mu subunits of the adaptor protein (AP) complex.[39] The signal-transducing adaptor protein 1 (STAP1) Src homology 2 (SH2) domain binding motif at Tyr446 - Ile450 (YELLI) is a phosphotyrosine (pTyr) pocket that serves as a docking site for the SH2 domain, which is central to tyrosine kinase signaling.[37][40] Multiple periodicities typical of an α-helix (periods of 3.6 residues in the hydrophobicity) encompass transmembrane domains.[41] 3 tandem repeats with core block lengths of 3 aa (GNYT, VSTF, STFI) are observed throughout the sequence.[33]

Secondary Structure[edit]

Helical Wheel of Transmembrane Domain 3.

Based on results by Ali2D, the secondary structure of SLC46A3 is rich in α-helices with random coils in between.[42] More precisely, the protein is predicted to be composed of 62.9% α-helix, 33.8% random coil, and 3.3% extended strand. The regions of α-helices span the majority of the transmembrane domains. The signal peptide is also predicted to form an α-helix, most likely in the h-region.[43] The amphipathic α-helices possess a particular orientation with charged/polar and nonpolar residues on opposite sides of the helix mainly due to the hydrophobic effect.[44]

Membrane Topology of SLC46A3.

Membrane topology of SLC46A3 shows the 11 α-helical transmembrane domains embedded in the membrane with the N-terminus oriented toward the extracellular region (or lumen of the ER) and the C-terminus extended to the cytoplasmic region.[45][46]

Tertiary Structure[edit]

Tertiary Structure of SLC46A3.
SLC46A3 Ligand Binding Sites. A: 78M, B: Y01, C: 37X.

Model for the tertiary structure of SLC46A3 was constructed by I-TASSER based on a homologous crystal structure of the human organic anion transporter MFSD10 (Tetran) with a TM-score of 0.853.[47][48][49] The structure contains a cluster of 17 α-helices that spans the membrane and random coils that connect those α-helices. Multiple ligand binding sites are also predicted to reside in the structure, including those for (2S)-2,3-dihydroxypropyl(7Z)-pentadec-7-enoate (78M), cholesterol hemisuccinate (Y01), and octyl glucose neopentyl glycol (37X).[50][51]

Ligand Binding Sites of SLC46A3[49]
Ligand C-score Cluster Size Ligand Binding Site Residues
78M 0.05 3 112, 116, 197, 198, 201, 204, 208
Y01 0.05 3 89, 241, 265, 269, 273, 391, 394, 399
37X 0.03 2 86, 89, 90, 94, 109, 136

Regulation of Gene Expression[edit]

Gene Level Regulation[edit]

Promoter[edit]

SLC46A3 carries 4 promoter regions that lead to different transcript variants as identified by ElDorado at Genomatix.[24] Promoter A supports transcript variant 1 (GXT_2836199).

SLC46A3 Promoters[24]
Promoter Name Start End Length (bp) Transcript
A GXP_190678 28718802 28720092 1291 GXT_2775378, GXT_29165870, GXT_23385588, GXT_2836199, GXT_26222267, GXT_22739111, GXT_23500299
B GXP_190676 28714934 28715973 1040 GXT_2785139
C GXP_190679 28713272 28714311 1040 GXT_2781051
D GXP_19677 28704518 28705557 1040 GXT_2781071

*The coordinates are for GRCh38.

Transcription Factors[edit]

Transcription factors (TFs) bind to the promoter region of SLC46A3 and modulate the transcription of the gene.[52] The table below shows a curated list of predicted TFs. MYC proto-oncogene (c-Myc), the strongest hit at Genomatix with a matrix similarity of 0.994, dimerizes with myc-associated factor X (MAX) to affect gene expression in a way that increases cell proliferation and cell metabolism.[53][54] Its expression is highly amplified in the majority of human cancers, including Burkitt's lymphoma. The heterodimer can repress gene expression by binding to myc-interacting zinc finger protein 1 (MIZ1), which also binds to the promoter of SLC46A3. CCAAT-displacement protein (CDP) and nuclear transcription factor Y (NF-Y) have multiple binding sites within the promoter sequence (3 sites for CDP and 2 sites for NF-Y).[53] CDP, also known as Cux1, is a transcriptional repressor.[55] NF-Y is a heterotrimeric complex of three different subunits (NF-YA, NF-YB, NF-YC) that regulates gene expression, both positively and negatively, by binding to the CCAAT box.[56]

SLC46A3 Transcription Factors[53]
Transcription Factor Description Matrix Similarity
HIF hypoxia inducible factor 0.989
c-Myc myelocytomatosis oncogene (c-Myc proto-oncogene) 0.994
GATA1 GATA-binding factor 1 0.983
PXR/RXR pregnane X receptor / retinoid X receptor heterodimer 0.833
RREB1 Ras-responsive element binding protein 1 0.815
TFCP2L1 transcription factor CP2-like 1 (LBP-9) 0.897
ZNF34 zinc finger protein 34 (KOX32) 0.852
MIZ1 myc-interacting zinc finger protein 1 (ZBTB17) 0.962
RFX5 regulatory factor X5 0.758
CEBPB CCAAT/enhancer-binding protein beta 0.959
KLF2 Kruppel-like factor 2 (LKLF) 0.986
CSRNP1 cysteine/serine-rich nuclear protein 1 (AXUD1) 1.000
CDP CCAAT-displacement protein (CDP/Cux) 0.983

0.949

0.955

NF-Y nuclear transcription factor Y 0.944

0.934

ZNF692 zinc finger protein 692 0.855
KAISO transcription factor Kaiso (ZBTB33) 0.991
SP4 transcription factor Sp4 0.908
ZBTB24 zinc finger and BTB domain containing 24 0.864
E2F4 E2F transcription factor 4 0.982

Expression Pattern[edit]

Gene Expression Array-Based Profile for SLC46A3.

RNAseq data show SLC46A3 most highly expressed in the liver, small intestine, and kidney and relatively low expression in the brain, skeletal muscle, salivary gland, placenta, and stomach.[18][19][57] In fetuses of 10 – 20 weeks, the adrenal gland and intestine report high expression while the heart, kidney, lung, and stomach demonstrate the opposite.[58] Microarray data from NCBI GEO present high expression in pancreatic islets, pituitary gland, lymph nodes, peripheral blood, and liver with percentile ranks of 75 or above.[59] Conversely, tissues among the most lowly expressed levels of SLC46A3 include bronchial epithelial cells, caudate nucleus, superior cervical ganglion, smooth muscle, and colorectal adenocarcinoma, all with percentile ranks below 15. Immunohistochemistry supports expression of the gene in the liver and kidney, as well as in skin tissues, while immunoblotting (western blotting) provides evidence for protein abundance in the liver and tonsils, in addition to in papilloma and glioma cells.[14]

In Situ Hybridization on Mouse Spinal Column and Cervical Spine. (a)-(c) spinal column of juvenile mouse (P4) and (d) cervical spine of adult mouse (P56).

In situ hybridization data show ubiquitous expression of the gene in mouse embryos at stage E14.5 and the adult mouse brain at postnatal days 56 (P56).[60][61] In the spinal column of juvenile mouse (P4), SLC46A3 is relatively highly expressed in the articular facet, neural arch, and anterior and posterior tubercles.[62] The dorsal horn shows considerable expression in the cervical spine of adult mouse (P56).[63]

Transcript Level Regulation[edit]

RNA-binding Proteins[edit]

RNA-binding proteins (RBPs) that bind to the 5' or 3' UTR regulate mRNA expression by getting involved in RNA processing and modification, nuclear export, localization, and translation.[64] A list of some of the most highly predicted RBPs in conserved regions of the 5' and 3' UTRs are shown below.

RNA-binding Proteins in 5' UTR[65]
Protein Description Motif P-value
MBNL1 (muscleblind like splicing regulator 1) modulates alternative splicing of pre-mRNAs; binds specifically to expanded dsCUG RNA with unusual size CUG repeats; contributes to myotonic dystrophy ygcuky 8.38×10−3

2.52×10−3

ZC3H10 (zinc finger CCCH-type containing 10) functions as a tumor suppressor by inhibiting the anchorage-independent growth of tumor cells; mitochondrial regulator ssagcgm 6.33×10−3
FXR2 (FMR1 autosomal homolog 2) associated with the 60S large ribosomal subunit of polyribosomes; may contribute to fragile X cognitive disability syndrome dgacrrr 7.01×10−3
SRSF7 (serine/arginine-rich splicing factor 7) critical for mRNA splicing as part of the spliceosome; involved in mRNA nuclear export and translation acgacg 6.44×10−3
FMR1 (FMRP translational regulator 1) associated with polyribosomes; involved in mRNA trafficking; negative regulator of translation kgacarg 7.53×10−3
HNRNPM (heterogenous nuclear ribonucleoprotein M) influences pre-mRNA processing, mRNA metabolism, and mRNA transport gguugguu 5.07×10−3
YBX2 (Y-box binding protein 2) regulates the stability and translation of germ cell mRNAs aacawcd 1.68×10−3
RBM24 (RNA binding motif protein 24) a tissue-specific splicing regulator; involved in mRNA stability wgwgugd 5.83×10−4
PABPC4 (poly(A) binding protein cytoplasmic 4) regulates stability of labile mRNA species in activated T cells; involved in translation in platelets and megakaryocytes aaaaaar 5.61×10−3
HuR (human antigen R) stabilizes mRNA by binding AU rich elements (AREs) uukruuu 4.61×10−3
RNA-binding Proteins in 3' UTR[65]
Protein Description Motif P-value
ENOX1 (ecto-NOX disulfide-thiol exchanger 1) involved in plasma membrane electron transport (PMET) pathways with alternating hydroquinone (NADH) oxidase and protein disulfide-thiol interchange activities hrkacag 5.17×10−4
CNOT4 (CCR4-NOT transcription complex subunit 4) subunit of CCR4-NOT complex; E3 ubiquitin ligase activity; interacts with CNOT1 gacaga 5.14×10−4
SRSF3 (serine/arginine-rich splicing factor 3) critical for mRNA splicing as part of the spliceosome; involved in mRNA nuclear export and translation wcwwc 4.00×10−4
KHDRBS2 (KH RNA binding domain containing, signal transduction associated 2) influences mRNA splice site selection and exon inclusion rauaaam 5.90×10−3
HuR (human antigen R) stabilizes mRNA by binding AREs uukruuu 7.12×10−3
RBMS3 (RNA-binding motif, single-stranded-interacting protein 3) (may be) involved in the control of RNA metabolism hauaua 1.89×10−3
KHDRBS1 (KH RNA binding domain containing, signal transduction associated 1) involved in alternative splicing, cell cycle regulation, RNA 3'-end formation, tumorigenesis, and regulation of human immunodeficiency virus (HIV) gene expression auaaaav 2.66×10−4
PABPN1 (poly(A) binding protein nuclear 1) binds to nascent poly(A) tails and directs polymerization of poly(A) tails at the 3' ends of eukaryotic transcripts araaga 9.11×10−3
RBM42 (RNA binding motif protein 42) involved in maintaining cellular ATP levels under stress by protecting target mRNAs aacuamg 4.44×10−4

miRNA[edit]

Several miRNAs have binding sites in the conserved regions of the 3' UTR of SLC46A3. The following miRNAs can negatively regulate the expression of the mRNA via RNA silencing.[66] Silencing mechanisms include mRNA cleavage and translation repression based on the level of complementarity between the miRNA and mRNA target sequences.

miRNAs[67][68]
Name Binding Site Sequence Target Score
hsa-miR-494-3p ATGTTTCA 97
hsa-miR-106b-5p GCACTTT – GCACTTT – GCACTTTA 94
hsa-miR-7159-5p TTGTTGA – TTGTTGAA 94
hsa-miR-5680 ATTTCTA – CATTTCT 91
hsa-miR-4477b TCCTTAAA – TCCTTAAA 91
hsa-miR-660-5p AATGGGT – AATGGGTA 89
hsa-miR-4319 CTCAGGGA 89
hsa-miR-7162-3p ACCTCAG 89
hsa-miR-137-3p AGCAATAA 88
hsa-miR-6071 CAGCAGAA 88
hsa-miR-597-3p GAGAACCA 86
hsa-miR-510-3p TTTCAAA – GTTTCAAA 86

Secondary Structure[edit]

Secondary Structure of 3' UTR.

The secondary structure of RNA holds both structural and functional significance.[69] Among various secondary structure motifs, the stem-loop structure (hairpin loop) is often conserved across species due to its role in RNA folding, protecting structural stability, and providing recognition sites for RBPs.[70] The 5' UTR region of SLC46A3 has 7 stem-loop structures identified and 3' UTR region a total of 10.[71] The majority of the binding sites of RBPs and miRNAs given above are located at a stem-loop structure, which is also true for the poly(A) signal at the 3' end.

Protein Level Regulation[edit]

Subcellular Localization[edit]

The k-Nearest Neighbor (k-NN) prediction by PSORTII predicts SLC46A3 to be mainly located at the plasma membrane (78.3%) and ER (17.4%), but also possibly at the mitochondrion (4.3%).[72] Immunofluorescent staining of SLC46A3 shows positivity in the plasma membrane, cytoplasm, and actin filaments, although positivity in the latter two is most likely due to the process of the protein being transported by myosin from the ER to the plasma membrane; myosin transports cargo-containing membrane vesicles along actin filaments.[14][73]

Post-Translational Modification[edit]

Conceptual Translation of SLC46A3.

The SLC46A3 protein contains a signal peptide that facilitates co-translational translocation and is cleaved between Thr20 and Gly21.[74][75] The resulting mature protein, 441 amino acids of length, is subject to further post-translational modifications (PTMs). The sequence has 3 N-glycosylation sites (Asn38, Asn46, Asn53), which are all located in the non-cytoplasmic region flanked by the signal peptide and the first transmembrane domain.[76] Ridigity of the N-terminal region close to the membrane is increased by O-GalNAc at Thr25.[77][78] O-GlcNAc at sites Ser227, Thr231, Ser445, and Ser459 are involved in the regulation of signaling pathways.[79][80] In fact, Ser445 and Ser459 are also subject to phosphorylation, where both sites are associated with casein kinase II (CKII), suggesting a crosstalking network that regulates protein activity.[81][82][83] Other highly conserved phosphorylation sites include Thr166, Ser233, Ser253, and Ser454, which are most likely targeted by kinases protein kinase C (PKC), CKII, PKC, and CKI/II, respectively. Conserved glycation sites at epsilon amino groups of lysines are predicted at Lys101, Lys239, and Lys374 with possible disrupting effects on molecular conformation and function of the protein.[84][85] S-palmitoylation, which help the protein bind more tightly to the membrane by contributing to protein hydrophobicity and membrane association, is predicted at Cys261 and Cys438.[86][87][88][89] S-palmitoylation can also modulate protein-protein interactions of SLC46A3 by changing the affinity of the protein for lipid rafts.

Homology and Evolution[edit]

Paralogs[edit]

SLC46A1: Also known as the proton-coupled folate transporter, SLC46A3 transports folate and antifolate substrates across cell membranes in a pH-dependent manner.[90]

SLC46A2: Aliases include thymic stromal cotransporter homolog, TSCOT, and Ly110. SLC46A2 is involved in symporter activity[91] and is a transporter of the immune second messenger 2'3'-cGAMP.[92]

SLC46A3 Paralogs[21][93]
Paralog Estimated Date of Divergence (MYA) Accession Number Sequence Length (aa) Sequence Identity (%) Sequence Similarity (%)
SLC46A1 724 NP_542400.2 459 31 49
SLC46A2 810 NP_149040.3 475 27 44

Orthologs[edit]

SLC46A3 is a highly conserved protein with orthologs as distant as fungi.[21][93] Closely related orthologs have been found in mammals with sequence similarities above 75% while moderately related orthologs come from species of birds, reptiles, amphibians, and fish with sequence similarities of 50-70%. More distantly related orthologs have sequence similarities below 50% and are invertebrates, placozoa, and fungi. The MFS, MFS_1, and transmembrane domains mostly remain conserved throughout species. A selected list of orthologs obtained through NCBI BLAST is shown in the table below.

SLC46A3 Orthologs[21][93][94]
Genus and Species Common Name Taxonomic Group Date of Divergence (MYA) Accession Number Sequence Length (aa) Sequence Identity (%) Sequence Similarity (%)
Homo sapiens Human Mammalia 0 NP_861450.1 461 100 100
Macaca mulatta Rhesus Monkey Mammalia 29 XP_014976295.2 460 95 96
Mus musculus House Mouse Mammalia 90 NP_001343931.1 460 75 86
Ornithorhynchus anatinus Platypus Mammalia 177 XP_028904425.1 462 68 81
Gallus gallus Chicken Aves 312 NP_001025999.1 464 51 69
Pseudonaja textilis Eastern Brown Snake Reptilia 312 XP_026564717.1 461 44 63
Xenopus tropicalis Tropical Clawed Frog Amphibia 352 XP_002934077.1 473 42 62
Danio rerio Zebrafish Actinopterygii 435 XP_021329877.1 463 42 62
Rhincodon typus Whale Shark Chondrichthyes 473 XP_020383213.1 456 39 56
Anneissia japonica Feather Star Crinoidea 684 XP_033118008.1 466 29 47
Pecten maximus Great Scallop Bivalvia 797 XP_033735180.1 517 24 40
Drosophila navojoa Fruit Fly Insecta 797 XP_030245348.1 595 19 34
Nematostella vectensis Starlet Sea Anemone Anthozoa 824 XP_001640625.1 509 28 46
Schmidtea mediterranea Flatworm Rhabditophora 824 AKN21695.1 483 23 38
Trichoplax adhaerens Trichoplax Tricoplacia 948 XP_002114167.1 474 19 36
Chytriomyces confervae C. confervae Chytridiomycetes 1105 TPX75507.1 498 23 40
Tuber magnatum White Truffle Pezizomycetes 1105 PWW79074.1 557 21 34
Cladophialophora bantiana C. bantiana Eurotiomycetes 1105 XP_016623985.1 587 21 32
Exophiala mesophila Black Yeast Eurotiomycetes 1105 RVX69813.1 593 19 32
Aspergillus terreus Mold Eurotiomycetes 1105 GES65939.1 604 19 31

Evolutionary History[edit]

Evolution Rate of SLC46A3.

The SLC46A3 gene first appeared in fungi approximately 1105 million years ago (MYA).[21] It evolves at a relatively moderate speed. A 1% change in the protein sequence requires about 6.2 million years. The SLC46A3 gene evolves about 4 times faster than cytochrome c and 2.5 times slower than fibrinogen alpha chain.

Function[edit]

As an MFS protein, SLC46A3 is a membrane transporter, mainly involved in the movement of substrates across the lipid bilayer.[9] The protein works via secondary active transport, where the energy for transport is provided by an electrochemical gradient.[95]

A proposed function of SLC46A3 of rising importance is the direct transport of maytansine-based catabolites from the lysosome to the cytoplasm by binding the macrolide structure of maytansine.[96] Among the different types of antibody-drug conjugates (ADCs), maytansine-based noncleavable linker ADC catabolites, such as lysine-MCC-DM1, are particularly responsive to SLC46A3 activity.[17] The protein functions independent of the cell surface target or cell line, thus is most likely to recognize maytansine or a moiety within the maytansine scaffold. Through transmembrane transport activity, the protein regulates catabolite concentration in the lysosome. In addition, SLC46A3 expression has been identified as a mechanism for resistance to ADCs with noncleavable maytansinoid and pyrrolobenzodiazepine warheads.[97] Although subcellular localization predictions have failed to identify the lysosome as a final destination of the protein, the YXXphi motif identified in the protein sequence has shown to direct lysosomal sorting.[39]

SLC46A3 may be involved in plasma membrane electron transport (PMET), a plasma membrane analog of the mitochondrial electron transport chain (ETC) that oxidizes intracellular NADH and contributes to aerobic energy production by supporting glycolytic ATP production.[98] The 3' UTR region of SLC46A3 includes a binding site for ENOX1, a protein highly involved in PMET.[65][99] The C-(X)2-C motif in the protein sequence also suggests possible oxidoreductase activity.[36]

Interacting Proteins[edit]

SLC46A3 has been found to generally interact with proteins involved in membrane transport, immune response, catalytic activity, or oxidation of substrates.[100] Some of the most definite and clinically important interactions include the following proteins.

Variants[edit]

SNPs are a very common type of genetic variation and are silent most of the time.[108] However, certain SNPs in the conserved or functionally important regions of the gene may have adverse effects on gene expression and function. Some of the SNPs with potentially damaging effects identified in the coding sequence of SLC46A3 are shown in the table below.

SNPs of SLC46A3[109]
SNP mRNA position Amino Acid Position Base Change Amino Acid Change Function Description
rs1456067444 554 1 [T/G] [M/R] missense start codon
rs749501877 679 46 [A/G] [N/S] missense N-glycosylation site
rs776889950 897 119 [T/G] [C/G] missense C-(X)2-C motif
rs1403613207 967 142 [G/A] [G/D] missense conserved substrate translocation pore
rs764198426 1322 261 [CT/-] [C/F] frameshift S-palmitoylation site
rs1373735793 1878 446 [T/C] [Y/H] missense YXXphi motif & STAP1 SH2 domain binding motif
rs1342327615 1906 455 [G/A] [S/N] missense phosphorylation & O-GlcNAc site
rs757225275

rs751982648

1917 459 [T/G]

[T/-]

[S/A]

[S/Q]

missense

frameshift

phosphorylation & O-GlcNAc site

f*The coordinates/positions are for GRCh38.p7.

Clinical Significance[edit]

Cancer/Tumor[edit]

The clinical significance of SLC46A3 surrounds the protein's activity as a transporter of maytansine-based ADC catabolites.[96] shRNA screens employing two libraries identified SLC46A3 as the only hit as a mediator of noncleavable maytansine-based ADC-dependent cytotoxicity, with q-values of 1.18×10−9 and 9.01×10−3.[17] Studies show either lost or significantly reduced SLC46A3 expression (-2.79 fold decrease by microarray with p-value 5.80×10−8) in T-DM1 (DM1 payload attached to antibody trastuzumab)-resistant breast cancer cells (KPL-4 TR).[11] In addition, siRNA knockdown in human breast tumor cell line BT-474M1 also results in resistance to T-DM1. Such association between loss of SLC46A3 expression and resistance to ADCs also applies to pyrrolobenzodiazepine warheads, signifying the important role of SLC46A3 in cancer treatment.[97]

CDP, one of SLC46A3's transcription factors, works as a tumor suppressor where CDP deficiency activates phosphoinositide 3-kinase (PI3K) signaling that leads to tumor growth.[110] The loss of heterozygosity and mutations of CDP are also associated with a variety of cancers.[111]

Prostate Cancer[edit]

Microarray analysis of SLC46A3 in two different prostate cancer cell lines, LNCaP (androgen-dependent) and DU145 (androgen-independent), show SLC46A3 expression in DU145 to be about 5 times as high as in LNCaP for percentile ranks and 1.5 times as high for transformed counts, demonstrating an association between SLC46A3 and accelerated cell growth of prostate cancer cells.[12] SLC46A3 possibly contributes to the androgen-independent manner of cancer development.

Hepatocellular Carcinoma (HCC)[edit]

SLC46A3 was found to be down-regulated in 83.2% of human HCC tissues based on western blot scores and qRT-PCR results on mRNA expression (p < 0.0001).[13] Overexpression of the gene also reduced resistance to sorafenib treatment and improved overall survival rate (p = 0.00085).

Papilloma & Glioma[edit]

Western blot analysis supports substantially strong expression of SLC46A3 in papilloma and glioma cells when compared to expression in the liver, one of the organs where the gene is most highly expressed.[14]

Obesity[edit]

A genome-wide association study on obesity identified 10 variants in the flanking 5′UTR region of SLC46A3 that were highly associated with diet fat (% energy) (p = 1.36×10−6 - 9.57×10−6).[15] In diet-induced obese (DIO) mice, SLC46A3 shows decreased gene expression following c-Jun N-terminal kinase 1 (JNK1) depletion, suggesting possible roles in insulin resistance as well as glucose/triglyceride homeostasis.[112]

SARS-CoV & SARS-CoV-2[edit]

Understanding the interaction between SLC46A3 and NSP2 in addition to the functions of each protein is critical to gaining insight into the pathogenesis of coronaviruses, namely SARS-CoV and SARS-CoV-2. The NSP2 protein domain resides in a region of the coronavirus replicase that is not particularly conserved across coronaviruses, and thus the altering protein sequence leads to significant changes in protein structure, leading to structural and functional variability.[106]

See also[edit]

References[edit]

  1. ^ a b c GRCh38: Ensembl release 89: ENSG00000139508Ensembl, May 2017
  2. ^ a b c GRCm38: Ensembl release 89: ENSMUSG00000029650Ensembl, 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. ^ a b c d e "SLC46A3". NCBI (National Center for Biotechnology Information) Gene.
  6. ^ a b "SLC46A3 Gene". GeneCards The Human Gene Database.
  7. ^ Nakai K, Horton P (2007). "Computational Prediction of Subcellular Localization". Protein Targeting Protocols. Methods in Molecular Biology. Vol. 390. Totowa, NJ: Humana Press. pp. 429–466. doi:10.1007/1-59745-466-4_29. ISBN 978-1-58829-702-0.
  8. ^ a b c "solute carrier family 46 member 3 isoform a precursor [Homo sapiens]". NCBI (National Center for Biotechnology Information) Protein.
  9. ^ a b "SLC46A3". OMIM (Online Mendelian Inheritance in Man).
  10. ^ a b "MFS". NCBI (National Center for Biotechnology Information) CDD (Conserved Domain Database).
  11. ^ a b Li G, Guo J, Shen BQ, Yadav DB, Sliwkowski MX, Crocker LM, et al. (July 2018). "Mechanisms of Acquired Resistance to Trastuzumab Emtansine in Breast Cancer Cells". Molecular Cancer Therapeutics. 17 (7): 1441–1453. doi:10.1158/1535-7163.mct-17-0296. PMID 29695635.
  12. ^ a b Kanaoka R, Kushiyama A, Seno Y, Nakatsu Y, Matsunaga Y, Fukushima T, et al. (2015-06-03). "Pin1 Inhibitor Juglone Exerts Anti-Oncogenic Effects on LNCaP and DU145 Cells despite the Patterns of Gene Regulation by Pin1 Differing between These Cell Lines". PLOS ONE. 10 (6): e0127467. Bibcode:2015PLoSO..1027467K. doi:10.1371/journal.pone.0127467. PMC 4454534. PMID 26039047.
  13. ^ a b Zhao Q, Zheng B, Meng S, Xu Y, Guo J, Chen LJ, et al. (June 2019). "Increased expression of SLC46A3 to oppose the progression of hepatocellular carcinoma and its effect on sorafenib therapy". Biomedicine & Pharmacotherapy. 114: 108864. doi:10.1016/j.biopha.2019.108864. PMID 30981107.
  14. ^ a b c d "SLC46A3 Polyclonal Antibody". ThermoFisher Scientific.
  15. ^ a b Comuzzie AG, Cole SA, Laston SL, Voruganti VS, Haack K, Gibbs RA, Butte NF (2012-12-14). "Novel genetic loci identified for the pathophysiology of childhood obesity in the Hispanic population". PLOS ONE. 7 (12): e51954. Bibcode:2012PLoSO...751954C. doi:10.1371/journal.pone.0051954. PMC 3522587. PMID 23251661.
  16. ^ a b Pfefferle S, Schöpf J, Kögl M, Friedel CC, Müller MA, Carbajo-Lozoya J, et al. (October 2011). "The SARS-coronavirus-host interactome: identification of cyclophilins as target for pan-coronavirus inhibitors". PLOS Pathogens. 7 (10): e1002331. doi:10.1371/journal.ppat.1002331. PMC 3203193. PMID 22046132.
  17. ^ a b c Hamblett KJ, Jacob AP, Gurgel JL, Tometsko ME, Rock BM, Patel SK, et al. (December 2015). "SLC46A3 Is Required to Transport Catabolites of Noncleavable Antibody Maytansine Conjugates from the Lysosome to the Cytoplasm". Cancer Research. 75 (24): 5329–40. doi:10.1158/0008-5472.can-15-1610. PMID 26631267.
  18. ^ a b c Fagerberg L, Hallström BM, Oksvold P, Kampf C, Djureinovic D, Odeberg J, et al. (February 2014). "Analysis of the human tissue-specific expression by genome-wide integration of transcriptomics and antibody-based proteomics". Molecular & Cellular Proteomics. 13 (2): 397–406. doi:10.1074/mcp.m113.035600. PMC 3916642. PMID 24309898.
  19. ^ a b c Duff MO, Olson S, Wei X, Garrett SC, Osman A, Bolisetty M, Plocik A, Celniker SE, Graveley BR (May 2015). "Genome-wide identification of zero nucleotide recursive splicing in Drosophila". Nature. 521 (7552): 376–9. Bibcode:2015Natur.521..376D. doi:10.1038/nature14475. PMC 4529404. PMID 25970244.
  20. ^ "SLC46A3". AceView.
  21. ^ a b c d e "BLAST: Basic Local Alignment Search Tool". NCBI (National Center for Biotechnology Information).
  22. ^ "Variation Viewer (GRCh38)". NCBI (National Center for Biotechnology Information).
  23. ^ "SLC46A3". PAXdb.
  24. ^ a b c "SLC46A3". Genomatix: ElDorado. Archived from the original on 2021-08-26. Retrieved 2020-08-01.
  25. ^ Pruitt K, Brown G, Tatusova T, Maglott D (2012-04-06). The Reference Sequence (RefSeq) Database. National Center for Biotechnology Information (US).
  26. ^ a b "Homo sapiens solute carrier family 46 member 3 (SLC46A3), transcript variant 1, mRNA". NCBI (National Center for Biotechnology Information) Nucleotide. 14 December 2020.
  27. ^ "Homo sapiens solute carrier family 46 member 3 (SLC46A3), transcript variant 2, mRNA". NCBI (National Center for Biotechnology Information) Nucleotide. 13 December 2020.
  28. ^ "Homo sapiens solute carrier family 46 member 3 (SLC46A3), transcript variant 3, mRNA". NCBI (National Center for Biotechnology Information) Nucleotide. 2 July 2020.
  29. ^ "PREDICTED: Homo sapiens solute carrier family 46 member 3 (SLC46A3), transcript variant X1, mRNA". NCBI (National Center for Biotechnology Information) Nucleotide. 16 May 2021.
  30. ^ a b c "solute carrier family 46 member 3 isoform a precursor [Homo sapiens]". NCBI (National Center for Biotechnology Information) Protein.
  31. ^ a b "solute carrier family 46 member 3 isoform b precursor [Homo sapiens]". NCBI (National Center for Biotechnology Information) Protein.
  32. ^ a b "solute carrier family 46 member 3 isoform X1 [Homo sapiens]". NCBI (National Center for Biotechnology Information) Protein.
  33. ^ a b c d e Brendel V, Bucher P, Nourbakhsh IR, Blaisdell BE, Karlin S (March 1992). "Methods and algorithms for statistical analysis of protein sequences". Proceedings of the National Academy of Sciences of the United States of America. 89 (6): 2002–6. Bibcode:1992PNAS...89.2002B. doi:10.1073/pnas.89.6.2002. PMC 48584. PMID 1549558.
  34. ^ Gasteiger E, Hoogland C, Gattiker A, Duvaud S, Wilkins MR, Appel RD, Bairoch A (2005), "Protein Identification and Analysis Tools on the ExPASy Server", The Proteomics Protocols Handbook, Totowa, NJ: Humana Press, pp. 571–607, doi:10.1385/1-59259-890-0:571, ISBN 978-1-58829-343-5
  35. ^ Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002). "Membrane Proteins". Molecular Biology of the Cell (4th ed.). Garland Science.
  36. ^ a b Miseta A, Csutora P (August 2000). "Relationship between the occurrence of cysteine in proteins and the complexity of organisms". Molecular Biology and Evolution. 17 (8): 1232–9. doi:10.1093/oxfordjournals.molbev.a026406. PMID 10908643.
  37. ^ a b Kumar M, Gouw M, Michael S, Sámano-Sánchez H, Pancsa R, Glavina J, et al. (January 2020). "ELM-the eukaryotic linear motif resource in 2020". Nucleic Acids Research. 48 (D1): D296–D306. doi:10.1093/nar/gkz1030. PMC 7145657. PMID 31680160.
  38. ^ "TRG_ENDOCYTIC_2". ELM (The Eukaryotic Linear Motif resource for Functional Sites in Proteins).
  39. ^ a b Pandey KN (October 2010). "Small peptide recognition sequence for intracellular sorting". Current Opinion in Biotechnology. 21 (5): 611–20. doi:10.1016/j.copbio.2010.08.007. PMC 2997389. PMID 20817434.
  40. ^ "LIG_SH2_STAP1". ELM (The Eukaryotic Linear Motif resource for Functional Sites in Proteins).
  41. ^ Eisenberg D, Weiss RM, Terwilliger TC (January 1984). "The hydrophobic moment detects periodicity in protein hydrophobicity". Proceedings of the National Academy of Sciences of the United States of America. 81 (1): 140–4. Bibcode:1984PNAS...81..140E. doi:10.1073/pnas.81.1.140. PMC 344626. PMID 6582470.
  42. ^ Zimmermann L, Stephens A, Nam SZ, Rau D, Kübler J, Lozajic M, et al. (July 2018). "A Completely Reimplemented MPI Bioinformatics Toolkit with a New HHpred Server at its Core". Journal of Molecular Biology. 430 (15): 2237–2243. doi:10.1016/j.jmb.2017.12.007. PMID 29258817. S2CID 22415932.
  43. ^ Reithmeier RA (1996). "Assembly of proteins into membranes". Biochemistry of Lipids, Lipoproteins and Membranes. New Comprehensive Biochemistry. Vol. 31. Elsevier. pp. 425–471. doi:10.1016/s0167-7306(08)60523-2. ISBN 978-0-444-82359-5.
  44. ^ Biggin PC, Sansom MS (February 1999). "Interactions of alpha-helices with lipid bilayers: a review of simulation studies". Biophysical Chemistry. 76 (3): 161–83. doi:10.1016/s0301-4622(98)00233-6. PMID 10074693.
  45. ^ Omasits U, Ahrens CH, Müller S, Wollscheid B (March 2014). "Protter: interactive protein feature visualization and integration with experimental proteomic data". Bioinformatics. 30 (6): 884–6. doi:10.1093/bioinformatics/btt607. hdl:20.500.11850/82692. PMID 24162465.
  46. ^ "Q7Z3Q1 (S46A3_HUMAN)". UniProt.
  47. ^ Yang J, Zhang Y (July 2015). "I-TASSER server: new development for protein structure and function predictions". Nucleic Acids Research. 43 (W1): W174-81. doi:10.1093/nar/gkv342. PMC 4489253. PMID 25883148.
  48. ^ Zhang Y, Skolnick J (2005-04-11). "TM-align: a protein structure alignment algorithm based on the TM-score". Nucleic Acids Research. 33 (7): 2302–9. doi:10.1093/nar/gki524. PMC 1084323. PMID 15849316.
  49. ^ a b "I-TASSER results". Zhang Lab.
  50. ^ Zhang C, Freddolino PL, Zhang Y (July 2017). "COFACTOR: improved protein function prediction by combining structure, sequence and protein-protein interaction information". Nucleic Acids Research. 45 (W1): W291–W299. doi:10.1093/nar/gkx366. PMC 5793808. PMID 28472402.
  51. ^ Yang J, Roy A, Zhang Y (October 2013). "Protein-ligand binding site recognition using complementary binding-specific substructure comparison and sequence profile alignment". Bioinformatics. 29 (20): 2588–95. doi:10.1093/bioinformatics/btt447. PMC 3789548. PMID 23975762.
  52. ^ Latchman DS (2004). "Methods for Studying Transcription Factors". Eukaryotic Transcription Factors. Vol. 270. Elsevier. pp. 23–53. doi:10.1016/b978-012437178-1/50008-4. ISBN 978-0-12-437178-1. PMC 1131717. PMID 2119171. {{cite book}}: |journal= ignored (help)
  53. ^ a b c "SLC46A3 Transcription Factor Binding Sites". Genomatix: MatInspector.
  54. ^ Miller DM, Thomas SD, Islam A, Muench D, Sedoris K (October 2012). "c-Myc and cancer metabolism". Clinical Cancer Research. 18 (20): 5546–53. doi:10.1158/1078-0432.CCR-12-0977. PMC 3505847. PMID 23071356.
  55. ^ Ellis T, Gambardella L, Horcher M, Tschanz S, Capol J, Bertram P, et al. (September 2001). "The transcriptional repressor CDP (Cutl1) is essential for epithelial cell differentiation of the lung and the hair follicle". Genes & Development. 15 (17): 2307–19. doi:10.1101/gad.200101. PMC 312776. PMID 11544187.
  56. ^ Wang GZ, Zhang W, Fang ZT, Zhang W, Yang MJ, Yang GW, et al. (July 2014). "Arsenic trioxide: marked suppression of tumor metastasis potential by inhibiting the transcription factor Twist in vivo and in vitro". Journal of Cancer Research and Clinical Oncology. 140 (7): 1125–36. doi:10.1007/s00432-014-1659-6. PMID 24756364. S2CID 6332740.
  57. ^ "Illumina bodyMap2 transcriptome". NCBI (National Center for Biotechnology Information) BioProject.
  58. ^ Szabo L, Morey R, Palpant NJ, Wang PL, Afari N, Jiang C, et al. (December 2016). "Erratum to: Statistically based splicing detection reveals neural enrichment and tissue-specific induction of circular RNA during human fetal development". Genome Biology. 17 (1): 263. doi:10.1186/s13059-016-1123-9. PMC 5165717. PMID 27993159.
  59. ^ Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, Block D, et al. (April 2004). "A gene atlas of the mouse and human protein-encoding transcriptomes". Proceedings of the National Academy of Sciences of the United States of America. 101 (16): 6062–7. Bibcode:2004PNAS..101.6062S. doi:10.1073/pnas.0400782101. PMC 395923. PMID 15075390.
  60. ^ "SLC46A3". GenePaint.
  61. ^ "SLC46A3 (Mouse Brain)". Allen Brain Atlas.
  62. ^ "Slc46a3 ISH: Mus musculus, Male, P4, variable". Allen Brain Atlas.
  63. ^ "Slc46a3 ISH: Mus musculus, Male, P56, variable". Allen Brain Atlas.
  64. ^ Brinegar AE, Cooper TA (September 2016). "Roles for RNA-binding proteins in development and disease". Brain Research. 1647: 1–8. doi:10.1016/j.brainres.2016.02.050. PMC 5003702. PMID 26972534.
  65. ^ a b c Paz I, Kosti I, Ares M, Cline M, Mandel-Gutfreund Y (July 2014). "RBPmap: a web server for mapping binding sites of RNA-binding proteins". Nucleic Acids Research. 42 (Web Server issue): W361-7. doi:10.1093/nar/gku406. PMC 4086114. PMID 24829458.
  66. ^ Macfarlane LA, Murphy PR (November 2010). "MicroRNA: Biogenesis, Function and Role in Cancer". Current Genomics. 11 (7): 537–61. doi:10.2174/138920210793175895. PMC 3048316. PMID 21532838.
  67. ^ Chen Y, Wang X (January 2020). "miRDB: an online database for prediction of functional microRNA targets". Nucleic Acids Research. 48 (D1): D127–D131. doi:10.1093/nar/gkz757. PMC 6943051. PMID 31504780.
  68. ^ "SLC46A3". miRDB.
  69. ^ Vandivier LE, Anderson SJ, Foley SW, Gregory BD (April 2016). "The Conservation and Function of RNA Secondary Structure in Plants". Annual Review of Plant Biology. 67 (1): 463–88. doi:10.1146/annurev-arplant-043015-111754. PMC 5125251. PMID 26865341.
  70. ^ Control of Messenger RNA Stability. 1993. doi:10.1016/c2009-0-03269-3. ISBN 9780120847822.
  71. ^ Zuker M (July 2003). "Mfold web server for nucleic acid folding and hybridization prediction". Nucleic Acids Research. 31 (13): 3406–15. doi:10.1093/nar/gkg595. PMC 169194. PMID 12824337.
  72. ^ Nakai K, Horton P (2007). "Computational Prediction of Subcellular Localization". Protein Targeting Protocols. Methods in Molecular Biology. Vol. 390. Totowa, NJ: Humana Press. pp. 429–466. doi:10.1007/1-59745-466-4_29. ISBN 978-1-58829-702-0.
  73. ^ "The Cell: A Molecular Approach. Sixth Edition. By Geoffrey M. Cooper and Robert E. Hausman. Sunderland (Massachusetts): Sinauer Associates. $142.95. xxv + 832 p.; ill.; index. [A Companion Website is available.] 2013". The Quarterly Review of Biology. 89 (4): 399. 2014. doi:10.1086/678645. ISBN 978-0-87893-964-0. ISSN 0033-5770.
  74. ^ Almagro Armenteros JJ, Tsirigos KD, Sønderby CK, Petersen TN, Winther O, Brunak S, et al. (April 2019). "SignalP 5.0 improves signal peptide predictions using deep neural networks" (PDF). Nature Biotechnology. 37 (4): 420–423. doi:10.1038/s41587-019-0036-z. PMID 30778233. S2CID 216678118.
  75. ^ Käll L, Krogh A, Sonnhammer EL (May 2004). "A combined transmembrane topology and signal peptide prediction method". Journal of Molecular Biology. 338 (5): 1027–36. doi:10.1016/j.jmb.2004.03.016. PMID 15111065.
  76. ^ Julenius K, Johansen MB, Zhang Y, Brunak S, Gupta R (2009). "Prediction of Glycosylation Sites in Proteins". Bioinformatics for Glycobiology and Glycomics. Chichester, UK: John Wiley & Sons, Ltd. pp. 163–192. doi:10.1002/9780470029619.ch9. ISBN 978-0-470-02961-9.
  77. ^ Steentoft C, Vakhrushev SY, Joshi HJ, Kong Y, Vester-Christensen MB, Schjoldager KT, et al. (May 2013). "Precision mapping of the human O-GalNAc glycoproteome through SimpleCell technology". The EMBO Journal. 32 (10): 1478–88. doi:10.1038/emboj.2013.79. PMC 3655468. PMID 23584533.
  78. ^ Essentials of glycobiology. Varki, Ajit (Third ed.). Cold Spring Harbor, New York. 2017. ISBN 978-1-62182-132-8. OCLC 960166742.{{cite book}}: CS1 maint: location missing publisher (link) CS1 maint: others (link)
  79. ^ Gupta R, Brunak S (2001). "Prediction of glycosylation across the human proteome and the correlation to protein function". Pacific Symposium on Biocomputing. Pacific Symposium on Biocomputing. WORLD SCIENTIFIC: 310–22. doi:10.1142/9789812799623_0029. ISBN 978-981-02-4777-5. PMID 11928486.
  80. ^ Fisi V, Miseta A, Nagy T (2017). "The Role of Stress-Induced O-GlcNAc Protein Modification in the Regulation of Membrane Transport". Oxidative Medicine and Cellular Longevity. 2017: 1308692. doi:10.1155/2017/1308692. PMC 5804373. PMID 29456783.
  81. ^ Wang C, Xu H, Lin S, Deng W, Zhou J, Zhang Y, et al. (February 2020). "GPS 5.0: An Update on the Prediction of Kinase-specific Phosphorylation Sites in Proteins". Genomics, Proteomics & Bioinformatics. 18 (1): 72–80. doi:10.1016/j.gpb.2020.01.001. PMC 7393560. PMID 32200042.
  82. ^ Blom N, Gammeltoft S, Brunak S (December 1999). "Sequence and structure-based prediction of eukaryotic protein phosphorylation sites". Journal of Molecular Biology. 294 (5): 1351–62. doi:10.1006/jmbi.1999.3310. PMID 10600390.
  83. ^ Blom N, Sicheritz-Pontén T, Gupta R, Gammeltoft S, Brunak S (June 2004). "Prediction of post-translational glycosylation and phosphorylation of proteins from the amino acid sequence". Proteomics. 4 (6): 1633–49. doi:10.1002/pmic.200300771. PMID 15174133. S2CID 18810164.
  84. ^ Johansen MB, Kiemer L, Brunak S (September 2006). "Analysis and prediction of mammalian protein glycation". Glycobiology. 16 (9): 844–53. doi:10.1093/glycob/cwl009. PMID 16762979.
  85. ^ Chen JH, Lin X, Bu C, Zhang X (2018-10-10). "Role of advanced glycation end products in mobility and considerations in possible dietary and nutritional intervention strategies". Nutrition & Metabolism. 15 (1): 72. doi:10.1186/s12986-018-0306-7. PMC 6180645. PMID 30337945.
  86. ^ Xie Y, Zheng Y, Li H, Luo X, He Z, Cao S, et al. (June 2016). "GPS-Lipid: a robust tool for the prediction of multiple lipid modification sites". Scientific Reports. 6 (1): 28249. Bibcode:2016NatSR...628249X. doi:10.1038/srep28249. PMC 4910163. PMID 27306108.
  87. ^ Aicart-Ramos C, Valero RA, Rodriguez-Crespo I (December 2011). "Protein palmitoylation and subcellular trafficking". Biochimica et Biophysica Acta (BBA) - Biomembranes. 1808 (12): 2981–94. doi:10.1016/j.bbamem.2011.07.009. PMID 21819967.
  88. ^ Ren J, Wen L, Gao X, Jin C, Xue Y, Yao X (November 2008). "CSS-Palm 2.0: an updated software for palmitoylation sites prediction". Protein Engineering, Design & Selection. 21 (11): 639–44. doi:10.1093/protein/gzn039. PMC 2569006. PMID 18753194.
  89. ^ Guan X, Fierke CA (December 2011). "Understanding Protein Palmitoylation: Biological Significance and Enzymology". Science China Chemistry. 54 (12): 1888–1897. doi:10.1007/s11426-011-4428-2. PMC 4240533. PMID 25419213.
  90. ^ "SLC46A1". NCBI (National Center for Biotechnology Information) Gene.
  91. ^ "SLC46A2". NCIB (National Center for Biotechnology Information) Gene.
  92. ^ Cordova, Anthony F.; Ritchie, Christopher; Böhnert, Volker; Li, Lingyin (June 23, 2021). "Human SLC46A2 Is the Dominant cGAMP Importer in Extracellular cGAMP-Sensing Macrophages and Monocytes". ACS Cent Sci. 7 (6): 1073–1088. doi:10.1021/acscentsci.1c00440. PMC 8228594. PMID 34235268.
  93. ^ a b c Needleman SB, Wunsch CD (March 1970). "A general method applicable to the search for similarities in the amino acid sequence of two proteins". Journal of Molecular Biology. 48 (3): 443–53. doi:10.1016/0022-2836(70)90057-4. PMID 5420325.
  94. ^ Kumar S, Stecher G, Suleski M, Hedges SB (July 2017). "TimeTree: A Resource for Timelines, Timetrees, and Divergence Times". Molecular Biology and Evolution. 34 (7): 1812–1819. doi:10.1093/molbev/msx116. PMID 28387841.
  95. ^ Pao SS, Paulsen IT, Saier MH (March 1998). "Major facilitator superfamily". Microbiology and Molecular Biology Reviews. 62 (1): 1–34. doi:10.1128/mmbr.62.1.1-34.1998. PMC 98904. PMID 9529885.
  96. ^ a b Bissa B, Beedle AM, Govindarajan R (November 2016). "Lysosomal solute carrier transporters gain momentum in research". Clinical Pharmacology and Therapeutics. 100 (5): 431–436. doi:10.1002/cpt.450. PMC 5056150. PMID 27530302.
  97. ^ a b Kinneer K, Meekin J, Tiberghien AC, Tai YT, Phipps S, Kiefer CM, et al. (December 2018). "SLC46A3 as a Potential Predictive Biomarker for Antibody-Drug Conjugates Bearing Noncleavable Linked Maytansinoid and Pyrrolobenzodiazepine Warheads". Clinical Cancer Research. 24 (24): 6570–6582. doi:10.1158/1078-0432.ccr-18-1300. PMID 30131388.
  98. ^ Herst PM, Berridge MV (December 2006). "Plasma membrane electron transport: a new target for cancer drug development". Current Molecular Medicine. 6 (8): 895–904. doi:10.2174/156652406779010777. PMID 17168740. Retrieved 2020-08-01.
  99. ^ "ENOX1 ecto-NOX disulfide-thiol exchanger 1 [ Homo sapiens (human) ]". NCBI (National Center for Biotechnology Information) Gene.
  100. ^ "Figure S6: Predicted secondary structure of CoV-RMEN using CFSSP:Chou and Fasman secondary structure prediction server". doi:10.7717/peerj.9572/supp-13. {{cite journal}}: Cite journal requires |journal= (help)
  101. ^ Luck K, Kim DK, Lambourne L, Spirohn K, Begg BE, Bian W, et al. (April 2020). "A reference map of the human binary protein interactome". Nature. 580 (7803): 402–408. Bibcode:2020Natur.580..402L. doi:10.1038/s41586-020-2188-x. PMC 7169983. PMID 32296183.
  102. ^ "CD79A CD79a molecule [ Homo sapiens (human) ]". NCBI (National Center for Biotechnology Information) Gene.
  103. ^ "P11912 (CD79A_HUMAN)". UniProt.
  104. ^ Huttlin EL, Ting L, Bruckner RJ, Gebreab F, Gygi MP, Szpyt J, et al. (July 2015). "The BioPlex Network: A Systematic Exploration of the Human Interactome". Cell. 162 (2): 425–440. doi:10.1016/j.cell.2015.06.043. PMC 4617211. PMID 26186194.
  105. ^ "LGALS3 galectin 3 [ Homo sapiens (human) ]". NCBI (National Center for Biotechnology Information) Gene.
  106. ^ a b Graham RL, Sims AC, Baric RS, Denison MR (2006). "The NSP2 Proteins of Mouse Hepatitis Virus and Sars Coronavirus are Dispensable for Viral Replication". The Nidoviruses. Advances in Experimental Medicine and Biology. Vol. 581. Boston, MA: Springer US. pp. 67–72. doi:10.1007/978-0-387-33012-9_10. ISBN 978-0-387-26202-4. PMC 7123188. PMID 17037506.
  107. ^ "Review for "Therapeutic uncertainties in people with cardiometabolic diseases and severe acute respiratory syndrome coronavirus 2 ( <scp>SARS-CoV</scp> -2 or <scp>COVID</scp> -19)"". 2020-04-07. doi:10.1111/dom.14062/v1/review3. S2CID 219115750. {{cite journal}}: Cite journal requires |journal= (help)
  108. ^ Shen LX, Basilion JP, Stanton VP (July 1999). "Single-nucleotide polymorphisms can cause different structural folds of mRNA". Proceedings of the National Academy of Sciences of the United States of America. 96 (14): 7871–6. Bibcode:1999PNAS...96.7871S. doi:10.1073/pnas.96.14.7871. PMC 22154. PMID 10393914.
  109. ^ "SNP linked to Gene (geneID:283537) Via Contig Annotation". NCBI (National Center for Biotechnology Information) dbSNP Short Genetic Variations.
  110. ^ Wong CC, Martincorena I, Rust AG, Rashid M, Alifrangis C, Alexandrov LB, et al. (January 2014). "Inactivating CUX1 mutations promote tumorigenesis". Nature Genetics. 46 (1): 33–8. doi:10.1038/ng.2846. PMC 3874239. PMID 24316979.
  111. ^ Liu N, Sun Q, Wan L, Wang X, Feng Y, Luo J, Wu H (2020-05-29). "CUX1, A Controversial Player in Tumor Development". Frontiers in Oncology. 10: 738. doi:10.3389/fonc.2020.00738. PMC 7272708. PMID 32547943.
  112. ^ Yang R, Wilcox DM, Haasch DL, Jung PM, Nguyen PT, Voorbach MJ, et al. (August 2007). "Liver-specific knockdown of JNK1 up-regulates proliferator-activated receptor gamma coactivator 1 beta and increases plasma triglyceride despite reduced glucose and insulin levels in diet-induced obese mice". The Journal of Biological Chemistry. 282 (31): 22765–74. doi:10.1074/jbc.m700790200. PMID 17550900.

Further reading[edit]

Leave a Reply