TCDB is operated by the Saier Lab Bioinformatics Group

2.A.7 The Drug/Metabolite Transporter (DMT) Superfamily

The DMT Superfamily consists of 44 recognized families, each, in general, with a characteristic function, size and topology (Jack et al., 2001; Västermark et al., 2011). These phylogenetic families will be presented and described below; references, when available, will be provided, and representative well-characterized proteins, when available, will be tabulated. Lolkema et al. (2008) have presented bioinformatic analysis of prokaryotic members of the DMT (DUF606) superfamily concerning evolution of the antiparallel arrangements of the two homologous 5TMS domains. The primordial SMR-type permeases may have resulted from duplication of a 2 TMS-encoding genetic element, which added one TMS to give five TMSs, and then duplicated to give 10 TMS proteins: 2 → 4 → 5 → 10 (TMSs), (Lam et al., 2011). One member of the SMR-2 family (2.A.7.22.2) is a lipid (isoprenoid) flippase (Yan et al., 2007; Contreras et al., 2010).  Most nucleotide-sugar transporter in the endoplasmic reticulum and Golgi of eukaryotic cells are members of the DMT superfamily (Song 2013). Nucleotide-sugar transport into the Mammalian Golgi has been reviewed (Maszczak-Seneczko et al. 2022). DMT porters have the DMT fold (Ferrada and Superti-Furga 2022).


2.A.7.1 The 4 TMS Small Multidrug Resistance (SMR) Family

SMR family pumps are prokaryotic transport systems consisting of homodimeric or heterodimeric structures (Chung and Saier, 2001; Bay et al., 2007; Bay and Turner 2009). The subunits of these systems are of 100-120 amino acid residues in length and span the membrane as α-helices four times. Functionally characterized members of the SMR family catalyze multidrug efflux driven drug:H+ antiport where the proton motive force provides the driving force for drug efflux. The drugs transported are generally cationic, and a simple cation antiport mechanism involving the conserved Glu-14 has been proposed (Yerushalmi and Schuldiner, 2000). This mechanism suggests a requisite, mutually exclusive occupancy of Glu-14, providing a simple explanation for coupling the movement of two positively charged molecules. One system (YdgEF of E. coli; TC# 2.A.7.1.8) is reported to confer resistance to anionic detergents (Nishino and Yamaguchi, 2001). See 2.A.7.1.1 for substrates transporter by SMR systems (Lucero et al. 2023).

The 3-D structure of the dimeric EmrE shows opposite orientation of the two subunits in the membrane (Chen et al., 2007). The first three transmembrane helices from each monomer surround the substrate binding chamber, whereas the fourth helices participate only in dimer formation. Selenomethionine markers clearly indicate an antiparallel orientation for the monomers, supporting a 'dual topology' model. On the basis of available structural data, a model for the proton-dependent drug efflux mechanism of EmrE was proposed. Interestingly, Nasie et al., (2010) suggest that EmrE can insert randomly in two orientations and can exhibit activity in both the parallel and antiparallel orientations. The orientation of small multidrug resistance transporter subunits in the membrane correlate with the positive-inside rule (Kolbusz et al., 2010). A comprehensive review of the classes of efflux pump inhibitors from various sources, highlighting their structure-activity relationships, which can be useful for medicinal chemists in the pursuit of novel efflux pump inhibitors, has appeared (Durães et al. 2018).  Drug metabolising enzymes and transporters in the paediatric duodenum have been quantitated (Goelen et al. 2023).  Members of the SMRGdx subtype can export the degradation products of metformin, helping bacteria adapt to high environmental levels of the commonly prescribed diabetes medication (Short 2024).


2.A.7.2 The 5 TMS Bacterial/Archaeal Transporter (BAT) Family

The BAT family consists of 5 TMS proteins from bacteria and archaea. None of these proteins is functionally characterized.


2.A.7.3 The 10 TMS Drug/Metabolite Exporter (DME) Family

The DME family is a large family of integral membrane proteins with sizes ranging from 287 to 310 amino acyl residues and exhibiting 10 putative α-helical transmembrane spanners (TMSs). These proteins are derived from phylogenetically divergent bacteria and archaea, and B. subtilis, E. coli, S. coelicolor and A. fulgidus have multiple paralogues. Distant eukaryotic homologues are more closely related to DME family members than to other DM superfamily members can be found (i.e., the Riken gene product of the mouse (BAC31006)).

Proteins of the DME family evidently arose by an internal gene duplication event as the first halves of these proteins are homologous to the second halves. One of these prokaryotic proteins, YdeD, is functionally characterized and exports cysteine metabolites in E. coli. Another, RhtA of E. coli, exports threonine and homoserine. A third, Sam of Rickettsia prowazekii, takes up S-adenosylmethionine (TC #3.A.7.3.7; Tucker et al., 2003). In addition, several members of the DME family have been implicated in solute transport. Thus, the MttP protein of the archaeon, Methanosarcina barkeri, may transport methylamine (Ferguson and Krzycki, 1997); MadN is encoded within the malonate utilization operon of Malonomonas rubra and may be an acetate efflux pump, and PecM is encoded within a locus of Erwinia chrysanthemi controlling pectinase, cellulase and blue pigment production and might export the pigment indigoidine, produced by gene products encoded in the pecM operon. The PecM protein has been shown experimentally to exhibit a 10 TMS topology (Rouanet and Nasser, 2001).


2.A.7.4 The Plant Drug/Metabolite Exporter (P-DME) Family

The P-DME (UMAMIT) family is a large subset of the DME family. All of these proteins are derived from plants, and they cluster loosely together on a phylogenetic tree that includes all members of the DME and P-DME families. All of these proteins are predicted by various methods to have 10 TMSs. If this suggestion proves to be correct, then the two halves of these proteins will have opposite orientation in the membrane. The P-DME family members have been called UMAMIT, and form large gene families in Arabidopsis (47 members) and rice (53 members). The few characterized members from Arabidopsis mediate amino acid export from the cytosol, with two of them shown to function as facilitators. They can be found in the plasma membrane (Ladwig et al 2012, Muller et al 2015, Besnard et al 2016, 2018) or the vacuolar membrane (Ranocha et al, 2010, Besnard et al 2018). They play multiple role in amino acid translocation between the organs of the plants, e.g. from leaves to seeds or to roots”.


2.A.7.5 The Glucose/Ribose Porter (GRP) Family

The glucose/ribose uptake (GRU) family includes two functionally characterized members, a glucose uptake permease of Staphylococcus xylosus, and a probable ribose uptake permease of Lactobacillus sakei. Both proteins probably function by H+ symport.


2.A.7.6 The L-Rhamnose Transporter (RhaT) Family

The RhaT family includes only 2 proteins, the rhamnose:H+ symporters of E. coli and Salmonella typhimurium, both of which have been functionally characterized. The RhaT proteins of both species are 344 aas long with 10 putative TMSs.


2.A.7.7 The Chloramphenicol-Sensitivity Protein (RarD) Family

No member of the RarD family is functionally characterized. Members of the family are from Gram-negative bacteria, Gram-positive bacteria and possibly archaea. They vary in size from 250-300 residues. They exhibit 10 TMSs.


2.A.7.8 The Caenorhabditis elegans ORF (CEO) Family

The CEO family is a small family of 6 paralogues encoded within the genome of C. elegans. None of these proteins is functionally characterized.


2.A.7.9 The Triose-phosphate Transporter (TPT) Family

Functionally characterized members of the former TPT family are derived from the inner envelope membranes of chloroplasts and nongreen plastids of plants. However, homologues are also present in yeast. Saccharomyces cerevisiae has three functionally uncharacterized TPT paralogues encoded within its genome. Under normal physiological conditions, chloroplast TPTs mediate a strict antiport of substrates, frequently exchanging an organic three carbon compound phosphate ester for inorganic phosphate (Pi). Normally, a triose-phosphate, 3-phosphoglycerate, or another phosphorylated C3 compound made in the chloroplast during photosynthesis, exits the organelle into the cytoplasm of the plant cell in exchange for Pi. These transporters are members of a subfamily, the TPT subfamily within the TPT family. Experiments with reconstituted translocators in artificial membranes indicate that transport can also occur by a channel-like uniport mechanism with up to 10-fold higher transport rates. Channel opening may be induced by a membrane potential of large magnitude and/or by high substrate concentrations. Nongreen plastid and chloroplast carriers, such as those from maize endosperm and root membranes, mediate transport of C3 compounds phosphorylated at carbon atom 2, particularly phosphoenolpyruvate, in exchange for Pi. These are the phosphoenolpyruvate:Pi antiporters (the PPT subfamily). Glucose-6-P has also been shown to be a substrate of some plastid translocators (the GPT subfamily). These three subfamilies of proteins (TPT, PPT and GPT) are divergent in sequence as well as substrate specificity, but their substrate specificities overlap.

Each TPT family protein consists of about 400-450 amino acyl residues with 5-8 putative transmembrane α-helical spanners TMSs). The actual number has been proposed to be 6 for the plant proteins as for mitochondrial carriers (TC# 2.A.29) and members of several other transporter families. However, proteins of the TPT family do not exhibit significant sequence similarity with the latter proteins, and there is no evidence for an internal repeat sequence. TPT proteins may exist as homodimers in the membrane.

The generalized reaction catalyzed by the proteins of the TPT family is:

organic phosphate ester (in) + Pi (out) ⇌ organic phosphate ester (out) + Pi (in).


2.A.7.10 The UDP-N-Acetylglucosamine:UMP Antiporter (UAA) Family

Nucleotide-sugar transporters (NSTs) are found in the Golgi apparatus and the endoplasmic reticulum of eukaryotic cells. Members of the family have been sequenced from yeast, protozoans and animals. Animals such as C. elegans possess many of these transporters. Humans have at least two closely related isoforms of the UDP-galactose:UMP exchange transporter.

NSTs generally appear to function by antiport mechanisms, exchanging a nucleotide-sugar for a nucleotide. Thus, CMP-sialic acid is exchanged for CMP; GDP-mannose is preferentially exchanged for GMP, and UDP-galactose and UDP-N-acetylglucosamine are exchanged for UMP (or possibly UDP). Other nucleotide sugars (e.g., GDP-fucose, UDP-xylose, UDP-glucose, UDP-N-acetylgalactosamine, etc.) may also be transported in exchange for various nucleotides, but their transporters have not been molecularly characterized. Each compound appears to be translocated by its own transport protein. Transport allows the compound, synthesized in the cytoplasm, to be exported to the lumen of the Golgi apparatus or the endoplasmic reticulum where it is used for the synthesis of glycoproteins and glycolipids. Comparable transport proteins exist for ATP which phosphorylates proteins, and phosphoadenosine phosphosulfate (PAPS) which is used as a percursor for protein sulfation. It is not known if these transport proteins are members of the DMT superfamily.

The sequenced NSTs are generally of about 320-340 amino acyl residues in length and exhibit 8-12 putative transmembrane α-helical spanners. An 8 TMS model has been presented by Kawakita et al. (1998) for the human UDP galactose transporter 1.

The generalized reaction catalyzed by NSTs is:

nucleotide-sugar (cytoplasm) + nucleotide (lumen) ⇌ nucleotide-sugar (lumen) + nucleotide (cytoplasm)


2.A.7.11 The UDP-Galactose:UMP Antiporter (UGA) Family

Nucleotide-sugar transporters (NSTs) are found in the Golgi apparatus and the endoplasmic reticulum of eukaryotic cells. Members of the family have been sequenced from yeast, protozoans and animals. Animals such as C. elegans possess many of these transporters. Humans have at least two closely related isoforms of the UDP-galactose:UMP exchange transporter.

NSTs generally appear to function by antiport mechanisms, exchanging a nucleotide-sugar for a nucleotide. Thus, CMP-sialic acid is exchanged for CMP; GDP-mannose is preferentially exchanged for GMP, and UDP-galactose and UDP-N-acetylglucosamine are exchanged for UMP (or possibly UDP). Other nucleotide sugars (e.g., GDP-fucose, UDP-xylose, UDP-glucose, UDP-N-acetylgalactosamine, etc.) may also be transported in exchange for various nucleotides, but their transporters have not been molecularly characterized. Each compound appears to be translocated by its own transport protein. Transport allows the compound, synthesized in the cytoplasm, to be exported to the lumen of the Golgi apparatus or the endoplasmic reticulum where it is used for the synthesis of glycoproteins and glycolipids. Comparable transport proteins exist for ATP which phosphorylates proteins, and phosphoadenosine phosphosulfate (PAPS) which is used as a percursor for protein sulfation. It is not known if these transport proteins are members of the DMT superfamily.

The sequenced NSTs are generally of about 320-340 amino acyl residues in length and exhibit 8-12 putative transmembrane α-helical spanners. An 8 TMS model has been presented by Kawakita et al. (1998) for the human UDP galactose transporter 1.

The generalized reaction catalyzed by NSTs is:

nucleotide-sugar (cytoplasm) + nucleotide (lumen) ⇌ nucleotide-sugar (lumen) + nucleotide (cytoplasm)


2.A.7.12 The CMP-Sialate:CMP Antiporter (CSA) Family

Nucleotide-sugar transporters (NSTs) are found in the Golgi apparatus and the endoplasmic reticulum of eukaryotic cells. Members of the family have been sequenced from yeast, protozoans and animals. Animals such as C. elegans possess many of these transporters. Humans have at least two closely related isoforms of the UDP-galactose:UMP exchange transporter.

NSTs generally appear to function by antiport mechanisms, exchanging a nucleotide-sugar for a nucleotide. Thus, CMP-sialic acid is exchanged for CMP; GDP-mannose is preferentially exchanged for GMP, and UDP-galactose and UDP-N-acetylglucosamine are exchanged for UMP (or possibly UDP). Other nucleotide sugars (e.g., GDP-fucose, UDP-xylose, UDP-glucose, UDP-N-acetylgalactosamine, etc.) may also be transported in exchange for various nucleotides, but their transporters have not been molecularly characterized. Each compound appears to be translocated by its own transport protein. Transport allows the compound, synthesized in the cytoplasm, to be exported to the lumen of the Golgi apparatus or the endoplasmic reticulum where it is used for the synthesis of glycoproteins and glycolipids. Comparable transport proteins exist for ATP which phosphorylates proteins, and phosphoadenosine phosphosulfate (PAPS) which is used as a percursor for protein sulfation. It is not known if these transport proteins are members of the DMT superfamily.

The sequenced NSTs are generally of about 320-340 amino acyl residues in length and exhibit 8-12 putative transmembrane α-helical spanners. An 8 TMS model has been presented by Kawakita et al. (1998) for the human UDP galactose transporter 1.

The generalized reaction catalyzed by NSTs is:

nucleotide-sugar (cytoplasm) + nucleotide (lumen) ⇌ nucleotide-sugar (lumen) + nucleotide (cytoplasm)


2.A.7.13 The GDP-Mannose:GMP Antiporter (GMA) Family

The yeast VRG4 protein, also called 'vanidate resistance protein', is a GDP-mannose transporter with the same size and topology as the other NSTs, but it shows very little sequence similarity with them. Only with the PSI-BLAST program with one iteration do these proteins exhibit apparent similarity. VRG4 is most similar to proteins in C. elegans, Leishmania donovani, Arabidopsis thaliana, and another S. cerevisiae protein reported to be of 249 aas (spP40027).


2.A.7.14 The Plant Organocation Permease (POP) Family

A single member of the POP family (AtPUP1) has been functionally characterized. It has been shown to transport adenine and cytosine with high affinity. Evidence concerning energy coupling suggested an H+ symport mechanism. Purine derivatives (e.g., hypoxanthine), phytohormones (e.g., zeatin and kinetin) and alkaloids (e.g., caffeine) proved to be competitive inhibitors suggesting that they may be transport substrates. In fact trans-zeatin (a cytokinin) has been shown to be taken up, probably by at least two systems (Cedzich et al. 2008). The order of inhibition of adenine uptake by a variety of purine derivatives, phytohormones and alkaloids was reported to be: adenine, kinetin, caffeine, cytosine, zeatin, hypoxanthine, cytidine, nicotine, kinetin riboside, adenosine, zeatin riboside and thymine (Williams and Miller, 2001). At least 15 members of this family have been sequenced from A. thaliana (Gillissen et al., 2000). Thus, AtPUP1 may be a broad specificity organocation transporter. Other family members have been reported to exhibit different affinities for nucleobases.

The generalized transport reaction probably catalyzed by AtPUP1 is:

Organocation (out) + H+ (out) → Organocation (in) + H+ (in)


2.A.7.15 The UDP-glucuronate/UDP-N-acetylgalactosamine Transporter (UGnT) Family


2.A.7.16 The GDP-fucose Transporter (GFT) Family


2.A.7.17 The Aromatic Amino Acid/Paraquat Exporter (ArAA/P-E) Family

The ArAA/PE family is a small family of proteobacterial proteins with 10 putative TMSs and sizes and sequences that most resemble the proteins of the DME family (2.A.7.3) within the DMT superfamily. One member of this family, YddG of E. coli and Salmonella typhimurium (<95% identical), have been functionally characterized (Santiviago et al., 2002; Doroshenko et al., 2007). They are efflux pumps for paraquat (methyl viologen) which is a hydrophilic, doubly charged, quaternary ammonium compound that can participate in a redox cycle that generates oxygen free radicals in the bacterial cell under aerobic conditions. YddG cannot pump out acriflavin, showing that it is fairly specific. It also exports aromatic amino acids. Therefore, it may not be a multidrug resistance pump. Paraquat resistance is also dependent on the major Salmonella porin, OmpD. Thus, YddG and OmpD are believed to function together in exporting paraquat to the external medium, but it is not known if this occurs in one or two steps (Santiviago et al., 2002).

The overall reaction catalyzed by YddG is:

Paraquat (in) → Paraquat (out).


2.A.7.18 The Choline Uptake Transporter (LicB-T) Family

A single functionally characterized secondary transporter, LicB of Haemophilus influenzae defines the LicB-T family (Fan et al., 2003). It has 292 aas and 10 putative TMSs.

LicB is a high-affinity choline permease that takes up choline under choline-limiting conditions. It is required for the use of exogenous choline for the synthesis of phosphorylcholine which is incorporated into the bacterium's lipopolysaccharide (LPS). It does not play a role in osmoprotection. Phosphorylcholine derivatized LPS contributes to H. influenzae's pathogenesis by mimicry of host cell molecules (Fan et al., 2003).

The overall reaction catalyzed by LicB is probably:

choline (out) + H+ (out) → choline (in) + H+ (in).


2.A.7.19 The Nucleobase Uptake Transporter (NBUT) Family

The allantoin permeases of Phaseolus vulgaris (French bean) and Arabidopsis thaliana have been shown to transport uracil and fluorouracil as well as allantoin (Schmidt et al., 2004). Arabidopsis has several paralogues. Distant homologues are present in Bacteroides thetaiotamicron (AAO77915) and Entamoeba histolyticia (EAL46705). These proteins have 10 putative TMSs and comprise a distinct family in the DMT superfamily.


2.A.7.20 The Chloroquine Resistance Transporter (PfCRT) Family

The Plasmodium falciparum chloroquine resistance protein (PfCRT) is a transporter as are its homologues in various species. In Plasmodium species it is localized to the intra-erythrocytic digestive vacuole. Mutations in this protein confer Verapamil-reversible chloroquine resistance to P. falciparum. The mutations in PfCRT give rise to increased compartment acidification. PfCRT-related changes in chloroquine response involve altered drug flux across the parasite degestive vacuole membrane. It has been concluded that PfCRT directly mediates efflux of chloroquine from the digrestive vacuole (Bray et al., 2005).

PfCRT is a 423 amino acyl protein with 10 putative TMSs, it probably catalyzes chloroquine quinine flux with H+ across the digestive vacuole membrane (Wellems, 2002). It is a member of the drug metabolite transporter (DMT) superfamily (TC #2.A.7) (Tran and Saier, 2004). In frog oocytes it has been reported to activate various endogenous transporters (Nessler et al., 2004). It transports a variety of qunoline drugs including quinine and quinidine. Mutations in TMSs 1, 4 and 9 alter drug specificity and determine levels of accumulation, suggesting that these TMSs play a role in substrate binding (Cooper et al., 2007). Chloroquine-resistance reversers are substrates for mutant PfCRTs (Lehane and Kirk, 2010).


2.A.7.21 The 5 TMS Bacterial/Archaeal Transporter-2 (BAT2) Family

The BAT2 family consists of 5 TMS proteins that resemble BAT family (2.A.7.2) proteins in size and topology, but show almost no sequence similarity with them.


2.A.7.22 The 4 TMS Small Multidrug Resistance-2 (SMR2) Family

The SMR2 family consists of 4 TMS proteins, most about 110-130 aas long, but some longer, that resemble the SMR family (2.A.7.1) proteins in size and topology. However, they show almost no sequence similarity. Not all of them have the conserved glutamate in TMS1. All close members of this family are from bacteria, but one distant member from Neurospora crassa has this domain N-terminal, fused to a CysT flavodoxin domain followed by a C-terminal radical SAM domain (Nicolet and Drennan, 2004). This protein (gi85104851) is reported to be 1061 aas long. Because this is the only protein in the database with this combination of fused domains, it could be artifactual. Another homologue from Frankia alni (419 aas; gi111220000) has a putative 9 TMS topology with a C-terminal 300 residue hydrophilic domain. Another protein, the TibA precursor glycoprotein adhesin/invasin of E. coli (336 aas; gi72166756) has 8 or 9 putative TMSs plus a C-terminal hydrophilic domain of nearly 100 residues. It may be distantly related to members of the DME family (2.A.7.3).


2.A.7.23 The Putative Tryptophan Efflux (Trp-E) Family

Expression of the Bacillus subtilis tryptophan biosynthetic genes trpEDCFBA and trpG, as well as a putative tryptophan transport gene (trpP), are regulated in response to tryptophan by the trp RNA-binding attenuation protein (TRAP). TRAP regulates expression of these genes by transcription attenuation and translation control mechanisms. TRAP and tryptophan also regulate translation of ycbK, a gene that encodes a protein of 312 aas and 10 TMSs, distantly related to members of the DMT superfamily (Yakhnin et al., 2006).


2.A.7.24 The Thiamine Pyrophosphate Transporter (TPPT) Family

This family includes a diverse group of proteins from all types of eukaryotes as well as prokaryotes. The only one with an assigned probable function is the Thi74 protein of yeast. These proteins have 10 TMSs in a 2 + 8 arrangement (possibly 2 + 4 + 4). No mechanistic details of the transport process are available.

The reaction believed to be catalyzed by Thi74 is:

TPP (out) → TPP (in).


2.A.7.25 The NIPA Mg2+ Uptake Permease (NIPA) Family

Mutations in the NIPA1(SPG6) gene of man, named for 'nonimprinted in Prader-Willi/Angelman' has been implicated in one form of autosomal dominant hereditary spastic paraplegia (HSP), a neurodegenerative disorder characterized by progressive lower limb spasticity and weakness. HSP comprises more than 30 genetic disorders whose predominant feature is a spastic gait. Mutations in at least six genes have been associated with autosomal dominant HSP including NIPA1(SPG6).

Reduced magnesium concentration enhances expression of NIPA1 suggesting a role in cellular magnesium metabolism. Indeed, NIPA1 mediates Mg2+ uptake that is electrogenic, voltage-dependent, and saturable with a Michaelis constant of 0.69 ± 0.21 mM when expressed in Xenopus oocytes (Goytain et al. 2007). Subcellular localization with immunofluorescence showed that endogenous NIPA1 protein associates with early endosomes and the cell surface in a variety of neuronal and epithelial cells. As expected of a magnesium-responsive gene, altered magnesium concentration leads to a redistribution between the endosomal compartment and the plasma membrane; high magnesium results in diminished cell surface NIPA1 whereas low magnesium leads to accumulation in early endosomes and recruitment to the plasma membrane. The mouse NIPA1 mutants, T39R and G100R, corresponding to the respective human mutants, showed a loss of function when expressed in oocytes and altered trafficking in transfected COS7 cells. NIPA1 seems to encode a Mg2+ transporter, and the loss of function of NIPA1(SPG6) due to abnormal trafficking of the mutated protein provides the basis of the HSP phenotype (Goytain et al. 2007).

NIPA has nine putative TMSs. Its mechanism of action is not known. It could be a channel or a carrier, and its energy dependency has not been studied. Homologues are found in a wide variety of animals, plants, and fungi. However, this family is clearly a member of the DMT superfamily (M. H. Saier, unpublished results).

The transport reaction catalyzed by NIPA is:

Mg2+ (out) → Mg2+ (in)


2.A.7.26 The 4 TMS Small Multidrug Resistance-3 (SMR3) Family

YnfA is a 108 aa E. coli protein with 4 established TMSs and both the N- and C-termini in the periplasm (Drew et al., 2002). Its homologues are found in a broad range of Gram-negative and Gram-positive bacteria as well as archaea and eukaryotes. The sizes of bacterial homologues range from 98 aas to 132 aas, with a few exceptions. Plant proteins can be as large as 197aas. The first two TMSs are homologous to the second two in these 4 TMS proteins. A Methanosarciniae mazei homologue of 94 aas and a Geobacillus kaustophilus homologue of 104 aas have only 2 TMSs with 30 residue extensions C- and N-terminal, respectively. No functional data are available for any of its homologues. This family is the YnfA UPF0060 family. 


2.A.7.27 The Ca2+ Homeostasis Protein (Csg2) Family


2.A.7.29 The Uncharacterized DMT-1 (U-DMT1) Family


2.A.7.30 The Uncharacterized DMT-2 (U-DMT2) Family


2.A.7.31 The Uncharacterized DMT-3 (U-DMT3) Family

References associated with 2.A.7 family:

and Quamme GA. (2010). Molecular identification of ancient and modern mammalian magnesium transporters. Am J Physiol Cell Physiol. 298(3):C407-29. 19940067
Abe, M., H. Hashimoto, and K. Yoda. (1999). Molecular characterization of Vig4/Vrg4 GDP-mannose transporter of the yeast Saccharomyces cerevisiae. FEBS Lett. 458: 309-312. 10570930
Abeijon, C., E.C. Mandon, and C.B. Hirschberg. (1997). Transporters of nucleotide sugars, nucleotide sulfate and ATP in the Golgi apparatus. Trends. Biochem. Sci. 22: 203-207. 9204706
Abeijon, C., P.W. Robbins, and C.B. Hirschberg. (1996). Molecular cloning of the Golgi apparatus uridine diphosphate-N-acetylglucosamine transporter from Kluyveromyces lactis. Proc. Natl. Acad. Sci. USA 93: 5963-5968. 8650202
Adewuyi, E.O., Y. Sapkota, International Endogene Consortium Iec, andMe Research Team, International Headache Genetics Consortium Ihgc, A. Auta, K. Yoshihara, M. Nyegaard, L.R. Griffiths, G.W. Montgomery, D.I. Chasman, and D.R. Nyholt. (2020). Shared Molecular Genetic Mechanisms Underlie Endometriosis and Migraine Comorbidity. Genes (Basel) 11:. 32121467
Afroz, S., A.M. El-Ganiny, D.A. Sanders, and S.G. Kaminskyj. (2011). Roles of the Aspergillus nidulans UDP-galactofuranose transporter, UgtA in hyphal morphogenesis, cell wall architecture, conidiation, and drug sensitivity. Fungal Genet Biol 48: 896-903. 21693196
Airich, L.G., I.S. Tsyrenzhapova, O.V. Vorontsova, A.V. Feofanov, V.G. Doroshenko, and S.V. Mashko. (2010). Membrane topology analysis of the Escherichia coli aromatic amino acid efflux protein YddG. J. Mol. Microbiol. Biotechnol. 19: 189-197. 21042032
Aké, F.M., P. Joyet, J. Deutscher, and E. Milohanic. (2011). Mutational analysis of glucose transport regulation and glucose-mediated virulence gene repression in Listeria monocytogenes. Mol. Microbiol. 81: 274-293. 21564334
Ament, C.E., S. Steinmann, K. Evert, G.M. Pes, S. Ribback, I. Gigante, E. Pizzuto, J.M. Banales, P.M. Rodrigues, P. Olaizola, H. Wang, G. Giannelli, X. Chen, M. Evert, and D.F. Calvisi. (2023). Aberrant fucosylation sustains the NOTCH and EGFR/NF-kB pathways and has a prognostic value in human intrahepatic cholangiocarcinoma. Hepatology. [Epub: Ahead of Print] 36789652
Andersen, P.K., L. Veng, H.R. Juul-Madsen, R.K. Vingborg, C. Bendixen, and B. Thomsen. (2007). Gene expression profiling, chromosome assignment and mutational analysis of the porcine Golgi-resident UDP-N-acetylglucosamine transporter SLC35A3. Mol. Membr. Biol. 24: 519-530. 17710655
Antony, H.A., N.S. Topno, S.N. Gummadi, D. Siva Sankar, R. Krishna, and S.C. Parija. (2018). In silico modeling of Plasmodium falciparum chloroquine resistance transporter protein and biochemical studies suggest its key contribution to chloroquine resistance. Acta Trop 189: 84-93. [Epub: Ahead of Print] 30308208
Ashikov, A., F. Routier, J. Fuhlrott, Y. Helmus, M. Wild, R. Gerardy-Schahn, and H. Bakker. (2005). The human solute carrier gene SLC35B4 encodes a bifunctional nucleotide sugar transporter with specificity for UDP-xylose and UDP-N-acetylglucosamine. J. Biol. Chem. 280: 27230-27235. 15911612
Baitsch, D., C. Sandu, R. Brandsch, and G.L. Igloi. (2001). Gene cluster on pAO1 of Arthrobacter nicotinovorans involved in degradation of the plant alkaloid nicotine: cloning, purification, and characterization of 2,6-dihydroxypyridine 3-hydroxylase. J. Bacteriol. 183: 5262-5267. 11514508
Bakker, H., F. Routier, S. Oelmann, W. Jordi, A. Lommen, R. Gerardy-Schahn, and D. Bosch. (2005). Molecular cloning of two Arabidopsis UDP-galactose transporters by complementation of a deficient Chinese hamster ovary cell line. Glycobiology 15: 193-201. 15456736
Basting, D., M. Lorch, I. Lehner, and C. Glaubitz. (2008). Transport cycle intermediate in small multidrug resistance protein is revealed by substrate fluorescence. FASEB J. 22: 365-373. 17873100
Bay, D.C. and R.J. Turner. (2009). Diversity and evolution of the small multidrug resistance protein family. BMC Evol Biol 9: 140. 19549332
Bay, D.C., K.L. Rommens, and R.J. Turner. (2008). Small multidrug resistance proteins: a multidrug transporter family that continues to grow. Biochim. Biophys. Acta. 1778: 1814-1838. 17942072
Beeler, T., K. Gable, C. Zhao, and T. Dunn. (1994). A novel protein, CSG2p, is required for Ca2+ regulation in Saccharomyces cerevisiae. J. Biol. Chem. 269: 7279-7284. 8125941
Berg, M., H. Hilbi, and P. Dimroth. (1997). Sequence of a gene cluster from Malonomonas rubra encoding components of the malonate decarboxylase Na+ pump and evidence for their function. Eur J Biochem 245: 103-115. 9128730
Berger, F., G.M. Gomez, C.P. Sanchez, B. Posch, G. Planelles, F. Sohraby, A. Nunes-Alves, and M. Lanzer. (2023). pH-dependence of the Plasmodium falciparum chloroquine resistance transporter is linked to the transport cycle. Nat Commun 14: 4234. 37454114
Berninsone, P., H.Y. Hwang, I. Zemtseva, H.R. Horvitz, and C.B. Hirschberg. (2001). SQV-7, a protein involved in Caenorhabditis elegans epithelial invagination and early embryogenesis, transports UDP-glucuronic acid, UDP-N- acetylgalactosamine, and UDP-galactose. Proc. Natl. Acad. Sci. USA 98: 3738-3743. 11259660
Berninsone, P., M. Eckhardt, R. Gerardy-Schahn, and C.B. Hirschberg. (1997). Functional expression of the murine Golgi CMP-sialic acid transporter in Saccharomyces cerevisiae. J. Biol. Chem. 272: 12616-12619. 9139716
Besnard, J., C. Zhao, J.C. Avice, S. Vitha, A. Hyodo, G. Pilot, and S. Okumoto. (2018). Arabidopsis UMAMIT24 and 25 are amino acid exporters involved in seed loading. J Exp Bot 69: 5221-5232. 30312461
Besnard, J., R. Pratelli, C. Zhao, U. Sonawala, E. Collakova, G. Pilot, and S. Okumoto. (2016). UMAMIT14 is an amino acid exporter involved in phloem unloading in Arabidopsis roots. J Exp Bot 67: 6385-6397. 27856708
Binet, R., R.E. Fernandez, D.J. Fisher, and A.T. Maurelli. (2011). Identification and characterization of the Chlamydia trachomatis L2 S-adenosylmethionine transporter. MBio 2: e51-5111. 21558433
Bray, P.G., R.E. Martin, L. Tilley, S.A. Ward, K. Kirk, and D.A. Fidock. (2005). Defining the role of PfCRT in Plasmodium falciparum chloroquine resistance. Mol. Microbiol. 56: 323-333. 15813727
Buppan, P., S. Seethamchai, N. Kuamsab, P. Harnyuttanakorn, C. Putaporntip, and S. Jongwutiwes. (2018). Multiple Novel Mutations in Chloroquine Resistance Transporter Gene during Implementation of Artemisinin Combination Therapy in Thailand. Am J Trop Med Hyg. [Epub: Ahead of Print] 30141388
Burska, U.L. and J.N. Fletcher. (2014). Two plasmid-encoded genes of enteropathogenic Escherichia coli strain K798 promote invasion and survival within HEp-2 cells. APMIS 122: 922-930. 24939568
Caffaro, C.E., C.B. Hirschberg, and P.M. Berninsone. (2006). Independent and simultaneous translocation of two substrates by a nucleotide sugar transporter. Proc. Natl. Acad. Sci. USA 103: 16176-16181. 17060606
Caffaro, C.E., K. Luhn, H. Bakker, D. Vestweber, J. Samuelson, P. Berninsone, and C.B. Hirschberg. (2008). A single Caenorhabditis elegans Golgi apparatus-type transporter of UDP-glucose, UDP-galactose, UDP-N-acetylglucosamine, and UDP-N-acetylgalactosamine. Biochemistry 47: 4337-4344. 18341292
Carruthers, M.D., B.H. Bellaire, and F.C. Minion. (2010). Exploring the response of Escherichia coli O157:H7 EDL933 within Acanthamoeba castellanii by genome-wide transcriptional profiling. FEMS Microbiol. Lett. 312: 15-23. 20831595
Castro, R., A.R. Neves, L.L. Fonseca, W.A. Pool, J. Kok, O.P. Kuipers, and H. Santos. (2009). Characterization of the individual glucose uptake systems of Lactococcus lactis: mannose-PTS, cellobiose-PTS and the novel GlcU permease. Mol. Microbiol. 71: 795-806. 19054326
Cedzich, A., H. Stransky, B. Schulz, and W.B. Frommer. (2008). Characterization of cytokinin and adenine transport in Arabidopsis cell cultures. Plant Physiol. 148: 1857-1867. 18835995
Chan, K.F., P. Zhang, and Z. Song. (2010). Identification of essential amino acid residues in the hydrophilic loop regions of the CMP-sialic acid transporter and UDP-galactose transporter. Glycobiology 20: 689-701. 20181793
Chanket, W., M. Pipatthana, A. Sangphukieo, P. Harnvoravongchai, S. Chankhamhaengdecha, T. Janvilisri, and M. Phanchana. (2024). The complete catalog of antimicrobial resistance secondary active transporters in : evolution and drug resistance perspective. Comput Struct Biotechnol J 23: 2358-2374. 38873647
Chen, Y.J., O. Pornillos, S. Lieu, C. Ma, A.P. Chen, and G. Chang. (2007). X-ray structure of EmrE supports dual topology model. Proc. Natl. Acad. Sci. USA 104: 18999-19004. 18024586
Chintala, S., J. Tan, R. Gautam, M.E. Rusiniak, X. Guo, W. Li, W.A. Gahl, M. Huizing, R.A. Spritz, S. Hutton, E.K. Novak, and R.T. Swank. (2007). The Slc35d3 gene, encoding an orphan nucleotide sugar transporter, regulates platelet-dense granules. Blood 109: 1533-1540. 17062724
Chung, Y.J. and M.H. Saier, Jr. (2001). SMR-type multidrug resistance pumps. Curr. Opin. Drug. Discov. Dev. 4: 237-245. 11378963
Chung, Y.J. and M.H. Saier, Jr. (2002). Overexpression of the Escherichia coli sugE gene confers resistance to a narrow range of quaternary ammonium compounds. J. Bacteriol. 184: 2543-2545. 11948170
Contreras, F.X., L. Sánchez-Magraner, A. Alonso, and F.M. Goñi. (2010). Transbilayer (flip-flop) lipid motion and lipid scrambling in membranes. FEBS Lett. 584: 1779-1786. 20043909
Cooper, R.A., K.D. Lane, B. Deng, J. Mu, J.J. Patel, T.E. Wellems, X. Su, and M.T. Ferdig. (2007). Mutations in transmembrane domains 1, 4 and 9 of the Plasmodium falciparum chloroquine resistance transporter alter susceptibility to chloroquine, quinine and quinidine. Mol. Microbiol. 63: 270-282. 17163969
Coppée, R., A. Sabbagh, and J. Clain. (2020). Structural and evolutionary analyses of the Plasmodium falciparum chloroquine resistance transporter. Sci Rep 10: 4842. 32179795
Damodaran, S. and L.C. Strader. (2019). Indole 3-Butyric Acid Metabolism and Transport in. Front Plant Sci 10: 851. 31333697
Dassler, T., T. Maier, C. Winterhalter, and A. Böck. (2000). Identification of a major facilitator protein from Escherichia coli involved in efflux of metabolites of the cysteine pathway. Mol. Microbiol. 36: 1101-1112. 10844694
Dastvan, R., A.W. Fischer, S. Mishra, J. Meiler, and H.S. Mchaourab. (2016). Protonation-dependent conformational dynamics of the multidrug transporter EmrE. Proc. Natl. Acad. Sci. USA 113: 1220-1225. 26787875
De Rossi, E., M. Branzoni, R. Cantoni, A. Milano, G. Riccardi, and O. Ciferri. (1998). mmr, a Mycobacterium tuberculosis gene conferring resistance to small cationic dyes and inhibitors. J. Bacteriol. 180: 6068-6071. 9811672
Dean, N., Y.B. Zhang, and J.B. Poster. (1997). The VRG4 gene is required for GDP-mannose transport into the lumen of the golgi in the yeast, Saccharomyces cerevisiae. J. Biol. Chem. 272: 31908-31914. 0
Defoor, E., M.B. Kryger, and J. Martinussen. (2007). The orotate transporter encoded by oroP from Lactococcus lactis is required for orotate utilization and has utility as a food-grade selectable marker. Microbiology 153: 3645-3659. 17975072
Denancé, N., P. Ranocha, N. Oria, X. Barlet, M.P. Rivière, K.A. Yadeta, L. Hoffmann, F. Perreau, G. Clément, A. Maia-Grondard, G.C. van den Berg, B. Savelli, S. Fournier, Y. Aubert, S. Pelletier, B.P. Thomma, A. Molina, L. Jouanin, Y. Marco, and D. Goffner. (2013). Arabidopsis wat1 (walls are thin1)-mediated resistance to the bacterial vascular pathogen, Ralstonia solanacearum, is accompanied by cross-regulation of salicylic acid and tryptophan metabolism. Plant J. 73: 225-239. 22978675
Descoteaux, A., Y. Luo, S.J. Turco, and S.M. Beverley. (1995). A specialized pathway affecting virulence glycoconjugates of Leishmania. Science 269: 1869-1872. 7569927
Di Fede, E., A. Peron, E.A. Colombo, C. Gervasini, and A. Vignoli. (2021). SLC35F1 as a candidate gene for neurodevelopmental disorders resembling Rett syndrome. Am J Med Genet A 185: 2238-2240. 33821533
Doroshenko, V., L. Airich, M. Vitushkina, A. Kolokolova, V. Livshits, and S. Mashko. (2007). YddG from Escherichia coli promotes export of aromatic amino acids. FEMS Microbiol. Lett. 275: 312-318. 17784858
Drew, D., D. Sjöstrand, J. Nilsson, T. Urbig, C.N. Chin, J.W. de Gier, and G. von Heijne. (2002). Rapid topology mapping of Escherichia coli inner-membrane proteins by prediction and PhoA/GFP fusion analysis. Proc. Natl. Acad. Sci. USA 99: 2690-2695. 11867724
Durães, F., M. Pinto, and E. Sousa. (2018). Medicinal Chemistry Updates on Bacterial Efflux Pump Modulators. Curr. Med. Chem. 25: 6030-6069. 29424299
Eckhardt, M., M. Mühlenhoff, A. Bethe, and R. Gerardy-Schahn. (1996). Expression cloning of the Golgi CMP-sialic acid transporter. Proc. Natl. Acad. Sci. USA 93: 7572-7576. 8755516
Elbaz, Y., T. Salomon, and S. Schuldiner. (2008). Identification of a glycine motif required for packing in EmrE, a multidrug transporter from Escherichia coli. J. Biol. Chem. 283: 12276-12283. 18321856
Engel, J., P.S. Schmalhorst, T. Dörk-Bousset, V. Ferrières, and F.H. Routier. (2009). A single UDP-galactofuranose transporter is required for galactofuranosylation in Aspergillus fumigatus. J. Biol. Chem. 284: 33859-33868. 19840949
Escudero, L., V. Mariscal, and E. Flores. (2015). Functional Dependence between Septal Protein SepJ from Anabaena sp. Strain PCC 7120 and an Amino Acid ABC-Type Uptake Transporter. J. Bacteriol. 197: 2721-2730. 26078444
Fabbro, D., C. Mio, F. Fogolari, and G. Damante. (2021). A novel de novo NIPA1 missense mutation associated to hereditary spastic paraplegia. J Hum Genet. [Epub: Ahead of Print] 34108639
Farenholtz, J., N. Artelt, A. Blumenthal, K. Endlich, H.K. Kroemer, N. Endlich, and O. von Bohlen Und Halbach. (2019). Expression of Slc35f1 in the murine brain. Cell Tissue Res 377: 167-176. 30868340
Ferguson, D.J. and J.A. Krzycki. (1997). Reconstruction of trimethylamine-dependent coenzyme M methylation with the trimethylamine corrinoid protein and the isozymes of methyltransferase II from Methanosarcina barkeri. J. Bacteriol. 179: 846-852. 9006042
Ferrada, E. and G. Superti-Furga. (2022). A structure and evolutionary-based classification of solute carriers. iScience 25: 105096. 36164651
Fiegler, H., J. Bassias, I. Jankovic, and R. Brückner. (1999). Identification of a gene in Staphylococcus xylosus encoding a novel glucose uptake protein. J. Bacteriol. 181: 4929-4936. 10438764
Fischer, K., B. Kammerer, M. Gutensohn, B. Arbinger, A. Weber, R.E. Häusler, and U.I. Flügge. (1997). A new class of plastidic phosphate translocators: a putative link between primary and secondary metabolism by the phosphoenolpyruvate/phosphate antiporter. Plant Cell 9: 453-462. 9090886
Fleishman, S.J., S.E. Harrington, A. Enosh, D. Halperin, C.G. Tate, and N. Ben-Tal. (2006). Quasi-symmetry in the cryo-EM structure of EmrE provides the key to modeling its transmembrane domain. J. Mol. Biol. 364: 54-67. 17005200
Flores, E., R. Pernil, A.M. Muro-Pastor, V. Mariscal, I. Maldener, S. Lechno-Yossef, Q. Fan, C.P. Wolk, and A. Herrero. (2007). Septum-localized protein required for filament integrity and diazotrophy in the heterocyst-forming cyanobacterium Anabaena sp. strain PCC 7120. J. Bacteriol. 189: 3884-3890. 17369306
Flügge, U.I. (1992). Reaction mechanism and asymmetric orientation of the reconstituted chloroplast phosphate translocator. Biochim. Biophys. Acta 1110: 112-118. 1390831
Flügge, U.I. (1995). Phosphate translocation in the regulation of photosynthesis. J. Exp. Bot. 46: 1317-1323. 0
Flügge, U.I. (1999). Phosphate translocators in plastids. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50: 27-45. 15012202
Flügge, U.I. and H.W. Heldt. (1991). Metabolite translocators of the chloroplast envelope. Ann. Rev. Plant Phys. Plant Mol. Biol. 42: 129-144. 0
Flügge, U.I., A. Weber, K. Fischer, B. Loddenkötter, and H. Wallmeier. (1992). Structure and function of the chloroplast triose phosphate/phosphate translocator. In N. Murata (ed.), Research in Photosynthesis, Vol. 3. Dordrecht: The Netherlands: Kluwer Academic Publishers, pp. 667-674. 0
Franke, I., A. Resch, T. Dassler, T. Maier, and A. Böck. (2003). YfiK from Escherichia coli promotes export of O-acetylserine and cysteine. J. Bacteriol. 185: 1161-1166. 12562784
Franke, S., G. Grass, C. Rensing, and D.H. Nies. (2003). Molecular analysis of the copper-transporting efflux system CusCFBA of Escherichia coli. J. Bacteriol. 185: 3804-3812. 12813074
Fu, J., W. Qin, Q. Tong, Z. Li, Y. Shao, Z. Liu, C. Liu, Z. Wang, and X. Xu. (2022). A novel DNA methylation-driver gene signature for long-term survival prediction of hepatitis-positive hepatocellular carcinoma patients. Cancer Med. [Epub: Ahead of Print] 35637633
Fuentes, D.E., C.A. Navarro, J.C. Tantaleán, M.A. Araya, C.P. Saavedra, J.M. Pérez, I.L. Calderón, P.A. Youderian, G.C. Mora, and C.C. Vásquez. (2005). The product of the qacC gene of Staphylococcus epidermidis CH mediates resistance to β-lactam antibiotics in gram-positive and gram-negative bacteria. Res. Microbiol. 156: 472-477. 15862444
Ganas, P. and R. Brandsch. (2009). Uptake of L-nicotine and of 6-hydroxy-L-nicotine by Arthrobacter nicotinovorans and by Escherichia coli is mediated by facilitated diffusion and not by passive diffusion or active transport. Microbiology 155: 1866-1877. 19443550
Ganas, P., M. Mihasan, G.L. Igloi, and R. Brandsch. (2007). A two-component small multidrug resistance pump functions as a metabolic valve during nicotine catabolism by Arthrobacter nicotinovorans. Microbiology 153: 1546-1555. 17464069
Gao, F., F.G. Wang, R.R. Lyu, F. Xue, J. Zhang, and R. Huo. (2018). SLC35E3 identified as a target of novel‑m1061‑5p via microRNA profiling of patients with cardiovascular disease. Mol Med Rep 17: 5159-5167. 29393345
Gao, X. and N. Dean. (2000). Distinct protein domains of the yeast golgi GDP-mannose transporter mediate oligomer assembly and export from the endoplasmic reticulum. J. Biol. Chem. 275: 17718-17727. 10748175
García Angulo, V.A., H.R. Bonomi, D.M. Posadas, M.I. Serer, A.G. Torres, &.#.1.9.3.;. Zorreguieta, and F.A. Goldbaum. (2013). Identification and characterization of RibN, a novel family of riboflavin transporters from Rhizobium leguminosarum and other proteobacteria. J. Bacteriol. 195: 4611-4619. 23935051
Garcia-Martin, C., L. Baldomà, J. Badía, and J. Aguilar. (1992). Nucleotide sequence of the rhaR-sodA interval specifying rhaT in Escherichia coli. J Gen Microbiol 138: 1109-1116. 1339463
Gillissen, B., L. Bürkle, B. André, C. Kühn, D. Rentsch, B. Brandl, and W.B. Frommer. (2000). A new family of high-affinity transporters for adenine, cytosine, and purine derivatives in Arabidopsis. Plant Cell 12: 291-300. 10662864
Girardi, E., A. César-Razquin, S. Lindinger, K. Papakostas, J. Konecka, J. Hemmerich, S. Kickinger, F. Kartnig, B. Gürtl, K. Klavins, V. Sedlyarov, A. Ingles-Prieto, G. Fiume, A. Koren, C.H. Lardeau, R. Kumaran Kandasamy, S. Kubicek, G.F. Ecker, and G. Superti-Furga. (2020). A widespread role for SLC transmembrane transporters in resistance to cytotoxic drugs. Nat Chem Biol 16: 469-478. 32152546
Goelen, J., G. Farrell, J. McGeehan, C.M. Titman, N. J W Rattray, T.N. Johnson, R.D. Horniblow, and H.K. Batchelor. (2023). Quantification of drug metabolising enzymes and transporter proteins in the paediatric duodenum via LC-MS/MS proteomics using a QconCAT technique. Eur J Pharm Biopharm 191: 68-77. 37625656
Goto, S., M. Taniguchi, M. Muraoka, H. Toyoda, Y. Sado, M. Kawakita, and S. Hayashi. (2001). UDP-sugar transporter implicated in glycosylation and processing of Notch. Nat. Cell Biol. 3: 816-822. 11533661
Goytain, A., R.M. Hines, A. El-Husseini, and G.A. Quamme. (2007). NIPA1(SPG6), the basis for autosomal dominant form of hereditary spastic paraplegia, encodes a functional Mg2+ transporter. J. Biol. Chem. 282: 8060-8068. 17166836
Guerin, A., A.S. Aziz, C. Mutch, J. Lewis, C.Y. Go, and S. Mercimek-Mahmutoglu. (2015). Pyridox(am)ine-5-Phosphate Oxidase Deficiency Treatable Cause of Neonatal Epileptic Encephalopathy With Burst Suppression: Case Report and Review of the Literature. J Child Neurol 30: 1218-1225. 25296925
Gutiérrez-Preciado, A., A.G. Torres, E. Merino, H.R. Bonomi, F.A. Goldbaum, and V.A. García-Angulo. (2015). Extensive Identification of Bacterial Riboflavin Transporters and Their Distribution across Bacterial Species. PLoS One 10: e0126124. 25938806
Gyimesi, G. and M.A. Hediger. (2022). Systematic in silico discovery of novel solute carrier-like proteins from proteomes. PLoS One 17: e0271062. 35901096
Hadley, B., T. Litfin, C.J. Day, T. Haselhorst, Y. Zhou, and J. Tiralongo. (2019). Nucleotide Sugar Transporter SLC35 Family Structure and Function. Comput Struct Biotechnol J 17: 1123-1134. 31462968
Handford, M.G., F. Sicilia, F. Brandizzi, J.H. Chung, and P. Dupree. (2004). Arabidopsis thaliana expresses multiple Golgi-localised nucleotide-sugar transporters related to GONST1. Mol. Genet. Genomics 272: 397-410. 15480787
Hao, G.J., Y.H. Ding, H. Wen, X.F. Li, W. Zhang, H.Y. Su, D.M. Liu, and N.L. Xie. (2017). Attenuation of deregulated miR-369-3p expression sensitizes non-small cell lung cancer cells to cisplatin via modulation of the nucleotide sugar transporter SLC35F5. Biochem. Biophys. Res. Commun. 488: 501-508. 28511796
He, G.X., C. Zhang, R.R. Crow, C. Thorpe, H. Chen, S. Kumar, T. Tsuchiya, and M.F. Varela. (2011). SugE, a New Member of the SMR Family of Transporters, Contributes to Antimicrobial Resistance in Enterobacter cloacae. Antimicrob. Agents Chemother. 55: 3954-3957. 21576447
He, L., K. Vasiliou, and D.W. Nebert. (2009). Analysis and update of the human solute carrier (SLC) gene superfamily. Hum Genomics 3: 195-206. 19164095
Hepburn, A.C., N. Lazzarini, R. Veeratterapillay, L. Wilson, J. Bacardit, and R. Heer. (2021). Identification of CNGB1 as a Predictor of Response to Neoadjuvant Chemotherapy in Muscle-Invasive Bladder Cancer. Cancers (Basel) 13:. 34359804
Higashi, K., H. Ishigure, R. Demizu, T. Uemura, K. Nishino, A. Yamaguchi, K. Kashiwagi, and K. Igarashi. (2008). Identification of a spermidine excretion protein complex (MdtJI) in Escherichia coli. J. Bacteriol. 190: 872-878. 18039771
Hiraoka S., T. Furuichi, G. Nishimura, S. Shibata, M. Yanagishita, D.L. Rimoin, A. Superti-Furga, P.G. Nikkels, M. Ogawa, K. Katsuyama, H. Toyoda, A. Kinoshita-Toyoda, N. Ishida , K. Isono, Y. Sanai, D.H. Cohn, H. KosekI, S. Ikegawa. (2007). Nucleotide-sugar transporter SLC35D1 is critical to chondroitin sulfate synthesis in cartilage and skeletal development in mouse and human. Nat Med. 13: 1363-1367. 17952091
Ho, A.M., B.J. Coombes, T.T.L. Nguyen, D. Liu, S.L. McElroy, B. Singh, M. Nassan, C.L. Colby, B.R. Larrabee, R.M. Weinshilboum, M.A. Frye, and J.M. Biernacka. (2020). Mood-Stabilizing Antiepileptic Treatment Response in Bipolar Disorder: A Genome-Wide Association Study. Clin Pharmacol Ther 108: 1233-1242. 32627186
Höglund, P.J., K.J. Nordström, H.B. Schiöth, and R. Fredriksson. (2011). The solute carrier families have a remarkably long evolutionary history with the majority of the human families present before divergence of Bilaterian species. Mol Biol Evol 28: 1531-1541. 21186191
Hori, H., H. Yoneyama, R. Tobe, T. Ando, E. Isogai, and R. Katsumata. (2011). Inducible L-alanine exporter encoded by the novel gene ygaW (alaE) in Escherichia coli. Appl. Environ. Microbiol. 77: 4027-4034. 21531828
Huo, W., Y. Wang, T. Chen, T. Cao, Y. Zhang, Z. Shi, and S. Hou. (2022). Triclosan activates c-Jun/miR-218-1-3p/SLC35C1 signaling to regulate cell viability, migration, invasion and inflammatory response of trophoblast cells in vitro. BMC Pregnancy Childbirth 22: 470. 35668364
Ishida N, Yoshioka S, Chiba Y, Takeuchi M, Kawakita M. (1999a). Molecular cloning and functional expression of the human Golgi UDP-N-acetylglucosamine transporter. J Biochem (Tokyo). 126(1):68-77. 10393322
Ishida N, Yoshioka S, Iida M, Sudo K, Miura N, Aoki K, Kawakita M. (1999b). Indispensability of transmembrane domains of Golgi UDP-galactose transporter as revealed by analysis of genetic defects in UDP-galactose transporter-deficient murine had-1 mutant cell lines and construction of deletion mutants. J Biochem (Tokyo). 126(6):1107-17. 10578063
Ishikawa, H.O., T. Ayukawa, M. Nakayama, S. Higashi, S. Kamiyama, S. Nishihara, K. Aoki, N. Ishida, Y. Sanai, and K. Matsuno. (2010). Two pathways for importing GDP-fucose into the endoplasmic reticulum lumen function redundantly in the O-fucosylation of Notch in Drosophila. J. Biol. Chem. 285: 4122-4129. 19948734
Jack, D.L., M.L. Storms, J.H. Tchieu, I.T. Paulsen, and M.H. Saier, Jr. (2000). A broad-specificity multidrug efflux pump requiring a pair of homologous SMR-type proteins. J. Bacteriol. 182: 2311-2313. 10735877
Jack, D.L., N.M. Yang, and M.H. Saier, Jr. (2001). The drug/metabolite transporter superfamily. Eur J Biochem 268: 3620-3639. 11432728
Jiang, T., J. Yang, H. Yang, W. Chen, K. Ji, Y. Xu, and L. Yu. (2022). SLC35B4 Stabilizes c-MYC Protein by O-GlcNAcylation in HCC. Front Pharmacol 13: 851089. 35308201
Julius, A., L. Laur, C. Schanzenbach, and D. Langosch. (2017). BLaTM 2.0, a Genetic Tool Revealing Preferred Antiparallel Interaction of Transmembrane Helix 4 of the Dual-Topology Protein EmrE. J. Mol. Biol. [Epub: Ahead of Print] 28432015
Kamel, M.G., N.T. Nam, N.H.B. Han, A.E. El-Shabouny, A.M. Makram, F.A. Abd-Elhay, T.N. Dang, N.L.T. Hieu, V.T.Q. Huong, T.H. Tung, K. Hirayama, and N.T. Huy. (2017). Post-dengue acute disseminated encephalomyelitis: A case report and meta-analysis. PLoS Negl Trop Dis 11: e0005715. 28665957
Kamiyama, S., N. Sasaki, E. Goda, K. Ui-Tei, K. Saigo, H. Narimatsu, Y. Jigami, R. Kannagi, T. Irimura, and S. Nishihara. (2006). Molecular cloning and characterization of a novel 3''-phosphoadenosine 5''-phosphosulfate transporter, PAPST2. J. Biol. Chem. 281: 10945-10953. 16492677
Kanjo, N., K. Nakahigashi, K. Oeda, and H. Inokuchi. (2001). Isolation and characterization of a cDNA from soybean and its homolog from Escherichia coli, which both complement the light sensitivity of Escherichia coli hemH mutant strain VS101. Genes Genet Syst 76: 327-334. 11817648
Kato, K., N. Shitan, T. Shoji, and T. Hashimoto. (2015). Tobacco NUP1 transports both tobacco alkaloids and vitamin B6. Phytochemistry 113: 33-40. 24947336
Kawakita, M., N. Ishida, N. Miura, G.H. Sun-Wada, and S. Yoshioka. (1998). Nucleotide sugar transporters: elucidation of their molecular identity and its implication for future studies. J Biochem 123: 777-785. 9562605
Kim, J., J.G. Kim, Y. Kang, J.Y. Jang, G.J. Jog, J.Y. Lim, S. Kim, H. Suga, T. Nagamatsu, and I. Hwang. (2004). Quorum sensing and the LysR-type transcriptional activator ToxR regulate toxoflavin biosynthesis and transport in Burkholderia glumae. Mol. Microbiol. 54: 921-934. 15522077
Klammt, C., F. Löhr, B. Schäfer, W. Haase, V. Dötsch, H. Rüterjans, C. Glaubitz, and F. Bernhard. (2004). High level cell-free expression and specific labeling of integral membrane proteins. Eur J Biochem 271: 568-580. 14728684
Knappe, S., U.I. Flügge, and K. Fischer. (2003). Analysis of the plastidic phosphate translocator gene family in Arabidopsis and identification of new phosphate translocator-homologous transporters, classified by their putative substrate-binding site. Plant Physiol. 131: 1178-1190. 12644669
Kolbusz, M.A., R. ter Horst, D.J. Slotboom, and J.S. Lolkema. (2010). Orientation of small multidrug resistance transporter subunits in the membrane: correlation with the positive-inside rule. J. Mol. Biol. 402: 127-138. 20643145
Korres, H. and N.K. Verma. (2004). Topological analysis of glucosyltransferase GtrV of Shigella flexneri by a dual reporter system and identification of a unique reentrant loop. J. Biol. Chem. 279: 22469-22476. 15028730
Krapivinsky, G., L. Krapivinsky, S.C. Stotz, Y. Manasian, and D.E. Clapham. (2011). POST, partner of stromal interaction molecule 1 (STIM1), targets STIM1 to multiple transporters. Proc. Natl. Acad. Sci. USA 108: 19234-19239. 22084111
Lühn, K., A. Laskowska, J. Pielage, C. Klämbt, U. Ipe, D. Vestweber, and M.K. Wild. (2004). Identification and molecular cloning of a functional GDP-fucose transporter in Drosophila melanogaster. Exp Cell Res 301: 242-250. 15530860
Laadhar, S., R. Ben Mansour, S. Marrakchi, N. Miled, M. Ennouri, J. Fischer, M.A. Kaddechi, H. Turki, and F. Fakhfakh. (2019). Identification of a novel missense mutation in NIPAL4 gene: First 3D model construction predicted its pathogenicity. Mol Genet Genomic Med e1104. [Epub: Ahead of Print] 31876100
Ladwig, F., M. Stahl, U. Ludewig, A.A. Hirner, U.Z. Hammes, R. Stadler, K. Harter, and W. Koch. (2012). Siliques are Red1 from Arabidopsis acts as a bidirectional amino acid transporter that is crucial for the amino acid homeostasis of siliques. Plant Physiol. 158: 1643-1655. 22312005
Lam, V.H., J.H. Lee, A. Silverio, H. Chan, K.M. Gomolplitinant, T.L. Povolotsky, E. Orlova, E.I. Sun, C.H. Welliver, and M.H. Saier, Jr. (2011). Pathways of transport protein evolution: recent advances. Biol Chem 392: 5-12. 21194372
Lee, Y., T. Nishizawa, M. Takemoto, K. Kumazaki, K. Yamashita, K. Hirata, A. Minoda, S. Nagatoishi, K. Tsumoto, R. Ishitani, and O. Nureki. (2017). Structure of the triose-phosphate/phosphate translocator reveals the basis of substrate specificity. Nat Plants. [Epub: Ahead of Print] 28970497
Lefèvre, C., B. Bouadjar, A. Karaduman, F. Jobard, S. Saker, M. Ozguc, M. Lathrop, J.F. Prud''homme, and J. Fischer. (2004). Mutations in ichthyin a new gene on chromosome 5q33 in a new form of autosomal recessive congenital ichthyosis. Hum Mol Genet 13: 2473-2482. 15317751
Lehane AM. and Kirk K. (2010). Efflux of a range of antimalarial drugs and 'chloroquine resistance reversers' from the digestive vacuole in malaria parasites with mutant PfCRT. Mol Microbiol. 77(4):1039-51. 20598081
Lehner, I., D. Basting, B. Meyer, W. Haase, T. Manolikas, C. Kaiser, M. Karas, and C. Glaubitz. (2008). The key residue for substrate transport (Glu14) in the EmrE dimer is asymmetric. J. Biol. Chem. 283: 3281-3288. 18042544
Lescano, C.I., C. Martini, C.A. González, and M. Desimone. (2016). Allantoin accumulation mediated by allantoinase downregulation and transport by Ureide Permease 5 confers salt stress tolerance to Arabidopsis plants. Plant Mol. Biol. 91: 581-595. 27209043
Lescano, I., M.F. Bogino, C. Martini, T.M. Tessi, C.A. González, K. Schumacher, and M. Desimone. (2019). Arabidopsis thaliana Ureide Permease 5 (AtUPS5) connects cell compartments involved in Ureide metabolism. Plant Physiol. [Epub: Ahead of Print] 31862838
Leuzzi, A., M.L. Di Martino, R. Campilongo, M. Falconi, M. Barbagallo, L. Marcocci, P. Pietrangeli, M. Casalino, M. Grossi, G. Micheli, B. Colonna, and G. Prosseda. (2015). Multifactor Regulation of the MdtJI Polyamine Transporter in Shigella. PLoS One 10: e0136744. 26313003
Li, D. and S. Mukhopadhyay. (2019). Functional analyses of the UDP-galactose transporter SLC35A2 using the binding of bacterial Shiga toxins as a novel activity assay. Glycobiology. [Epub: Ahead of Print] 30834435
Li, J., A. Sae Her, and N.J. Traaseth. (2021). Asymmetric protonation of glutamate residues drives a preferred transport pathway in EmrE. Proc. Natl. Acad. Sci. USA 118:. 34607959
Li, W., K. Wu, Y. Liu, Y. Yang, W. Wang, X. Li, Y. Zhang, Q. Zhang, R. Zhou, and H. Tang. (2020). Molecular cloning of SLC35D3 and analysis of its role during porcine intramuscular preadipocyte differentiation. BMC Genet 21: 20. 32087688
Lim, L., C.P. Sayers, C.D. Goodman, and G.I. McFadden. (2016). Targeting of a Transporter to the Outer Apicoplast Membrane in the Human Malaria Parasite Plasmodium falciparum. PLoS One 11: e0159603. 27442138
Linka, M., A. Jamai, and A.P. Weber. (2008). Functional characterization of the plastidic phosphate translocator gene family from the thermo-acidophilic red alga Galdieria sulphuraria reveals specific adaptations of primary carbon partitioning in green plants and red algae. Plant Physiol. 148: 1487-1496. 18799657
Lipiński, P., K.M. Stępień, E. Ciara, A. Tylki-Szymańska, and A. Jezela-Stanek. (2021). Skeletal and Bone Mineral Density Features, Genetic Profile in Congenital Disorders of Glycosylation: Review. Diagnostics (Basel) 11:. 34441372
Livshits, V.A., N.P. Zakataeva, B.B. Aleshin, and M.V. Vitushkina. (2003). Identification and characterization of the new gene rhtA involved in threonine and homoserine efflux in Escherichia coli. Res. Microbiol. 154: 123-135. 12648727
Lloris-Garcerá, P., J.S. Slusky, S. Seppälä, M. Prieß, L.V. Schäfer, and G. von Heijne. (2013). In vivo trp scanning of the small multidrug resistance protein EmrE confirms 3D structure models'. J. Mol. Biol. 425: 4642-4651. 23920359
Loddenkötter, B., B. Kammerer, K. Fischer, and U.I. Flügge. (1993). Expression of the functional mature chloroplast triose phosphate translocator in yeast internal membranes and purification of the histidine-tagged protein by a single metal-affinity chromatography step. Proc. Natl. Acad. Sci. USA 90: 2155-2159. 11607374
Lolkema, J.S., A. Dobrowolski, and D.J. Slotboom. (2008). Evolution of antiparallel two-domain membrane proteins: tracing multiple gene duplication events in the DUF606 family. J. Mol. Biol. 378: 596-606. 18384811
Lu, L., S. Varshney, Y. Yuan, H.X. Wei, A. Tanwar, S. Sundaram, M. Nauman, R.S. Haltiwanger, and P. Stanley. (2023). In vivo evidence for GDP-fucose transport in the absence of transporter SLC35C1 and putative transporter SLC35C2. J. Biol. Chem. 299: 105406. 38270391
Lu, L., X. Hou, S. Shi, C. Körner, and P. Stanley. (2010). Slc35c2 promotes Notch1 fucosylation and is required for optimal Notch signaling in mammalian cells. J. Biol. Chem. 285: 36245-36254. 20837470
Lu, S., X. Sun, H. Tang, J. Yu, B. Wang, R. Xiao, J. Qu, F. Sun, Z. Deng, C. Li, P. Yang, Z. Yang, and B. Rao. (2024). Colorectal cancer with low SLC35A3 is associated with immune infiltrates and poor prognosis. Sci Rep 14: 329. 38172565
Lucero, R.M., K. Demirer, T.J. Yeh, and R.B. Stockbridge. (2023). Transport of metformin metabolites by guanidinium exporters of the Small Multidrug Resistance family. bioRxiv. 37645731
Luck, K., D.K. Kim, L. Lambourne, K. Spirohn, B.E. Begg, W. Bian, R. Brignall, T. Cafarelli, F.J. Campos-Laborie, B. Charloteaux, D. Choi, A.G. Coté, M. Daley, S. Deimling, A. Desbuleux, A. Dricot, M. Gebbia, M.F. Hardy, N. Kishore, J.J. Knapp, I.A. Kovács, I. Lemmens, M.W. Mee, J.C. Mellor, C. Pollis, C. Pons, A.D. Richardson, S. Schlabach, B. Teeking, A. Yadav, M. Babor, D. Balcha, O. Basha, C. Bowman-Colin, S.F. Chin, S.G. Choi, C. Colabella, G. Coppin, C. D''Amata, D. De Ridder, S. De Rouck, M. Duran-Frigola, H. Ennajdaoui, F. Goebels, L. Goehring, A. Gopal, G. Haddad, E. Hatchi, M. Helmy, Y. Jacob, Y. Kassa, S. Landini, R. Li, N. van Lieshout, A. MacWilliams, D. Markey, J.N. Paulson, S. Rangarajan, J. Rasla, A. Rayhan, T. Rolland, A. San-Miguel, Y. Shen, D. Sheykhkarimli, G.M. Sheynkman, E. Simonovsky, M. Taşan, A. Tejeda, V. Tropepe, J.C. Twizere, Y. Wang, R.J. Weatheritt, J. Weile, Y. Xia, X. Yang, E. Yeger-Lotem, Q. Zhong, P. Aloy, G.D. Bader, J. De Las Rivas, S. Gaudet, T. Hao, J. Rak, J. Tavernier, D.E. Hill, M. Vidal, F.P. Roth, and M.A. Calderwood. (2020). A reference map of the human binary protein interactome. Nature 580: 402-408. 32296183
Lüders, F., H. Segawa, D. Stein, E.M. Selva, N. Perrimon, S.J. Turco, and U. Häcker. (2003). slalom encodes an adenosine 3'-phosphate 5'-phosphosulfate transporter essential for development in Drosophila. EMBO J. 22: 3635-3644. 12853478
Luhn, K., M.K. Wild, M. Eckhardt, R. Gerardy-Schahn, and D. Vestweber. (2001). The gene defective in leukocyte adhesion deficiency II encodes a putative GDP-fucose transporter. Nat. Genet. 28: 69-72. 11326279
Lytvynenko, I., S. Brill, C. Oswald, and K.M. Pos. (2016). Molecular basis of polyspecificity of the Small Multidrug Resistance Efflux Pump AbeS from Acinetobacter baumannii. J. Mol. Biol. 428: 644-657. 26707198
Ma, D., D.G. Russell, S.M. Beverley, and S.J. Turco. (1997). Golgi GDP-mannose uptake requires Leishmania LPG2: A member of a eukaryotic family of putative nucleotide-sugar transporters. J. Biol Chem. 272: 3799-3805. 9013638
Maeda, F., A. Kato, K. Takeshima, M. Shibazaki, R. Sato, T. Shibata, K. Miyake, H. Kozuka-Hata, M. Oyama, E. Shimizu, S. Imoto, S. Miyano, S. Adachi, T. Natsume, K. Takeuchi, Y. Maruzuru, N. Koyanagi, A. Jun, and K. Yasushi. (2022). Role of the Orphan Transporter SLC35E1 in the Nuclear Egress of Herpes Simplex Virus 1. J. Virol. 96: e0030622. 35475666
Maouche, R., H.L. Burgos, L. My, J.P. Viala, R.L. Gourse, and E. Bouveret. (2016). Coexpression of Escherichia coli obgE, Encoding the Evolutionarily Conserved Obg GTPase, with Ribosomal Proteins L21 and L27. J. Bacteriol. 198: 1857-1867. 27137500
Marchocki, Z., A. Tone, C. Virtanen, R. de Borja, B. Clarke, T. Brown, and T. May. (2022). Impact of neoadjuvant chemotherapy on somatic mutation status in high-grade serous ovarian carcinoma. J Ovarian Res 15: 50. 35501919
Marín-Quílez, A., L. Díaz-Ajenjo, C.A. Di Buduo, A. Zamora-Cánovas, M.L. Lozano, R. Benito, J.R. González-Porras, A. Balduini, J. Rivera, and J.M. Bastida. (2023). Inherited Thrombocytopenia Caused by Variants in Crucial Genes for Glycosylation. Int J Mol Sci 24:. 36982178
Martin, R.E., R.V. Marchetti, A.I. Cowan, S.M. Howitt, S. Bröer, and K. Kirk. (2009). Chloroquine transport via the malaria parasite's chloroquine resistance transporter. Science 325: 1680-1682. 19779197
Masaoka, Y., Y. Ueno, Y. Morita, T. Kuroda, T. Mizushima, and T. Tsuchiya. (2000). A two-component multidrug efflux pump, EbrAB, in Bacillus subtilis. J. Bacteriol. 182: 2307-2310. 10735876
Maszczak-Seneczko, D., M. Wiktor, E. Skurska, W. Wiertelak, and M. Olczak. (2022). Delivery of Nucleotide Sugars to the Mammalian Golgi: A Very Well (un)Explained Story. Int J Mol Sci 23:. 35955785
Maughan, S.C., M. Pasternak, N. Cairns, G. Kiddle, T. Brach, R. Jarvis, F. Haas, J. Nieuwland, B. Lim, C. Müller, E. Salcedo-Sora, C. Kruse, M. Orsel, R. Hell, A.J. Miller, P. Bray, C.H. Foyer, J.A. Murray, A.J. Meyer, and C.S. Cobbett. (2010). Plant homologs of the Plasmodium falciparum chloroquine-resistance transporter, PfCRT, are required for glutathione homeostasis and stress responses. Proc. Natl. Acad. Sci. USA 107: 2331-2336. 20080670
Mégret-Cavalier, M., A. Pozza, Q. Cece, F. Bonneté, I. Broutin, and G. Phan. (2024). Starting with an Integral Membrane Protein Project for Structural Biology: Production, Purification, Detergent Quantification, and Buffer Optimization-Case Study of the Exporter CntI from Pseudomonas aeruginosa. Methods Mol Biol 2715: 415-430. 37930543
Meier, A., H. Erler, and E. Beitz. (2018). Targeting Channels and Transporters in Protozoan Parasite Infections. Front Chem 6: 88. 29637069
Mermans, D., F. Nicolaus, K. Fleisch, and G. von Heijne. (2022). Cotranslational folding and assembly of the dimeric inner membrane protein EmrE. Proc. Natl. Acad. Sci. USA 119: e2205810119. 35994672
Mo, X.B., Y.H. Sun, Y.H. Zhang, and S.F. Lei. (2020). Integrative analysis highlighted susceptibility genes for rheumatoid arthritis. Int Immunopharmacol 86: 106716. 32599322
Mohamed, M., A. Ashikov, M. Guillard, J.H. Robben, S. Schmidt, B. van den Heuvel, A.P. de Brouwer, R. Gerardy-Schahn, P.M. Deen, R.A. Wevers, D.J. Lefeber, and E. Morava. (2013). Intellectual disability and bleeding diathesis due to deficient CMP-sialic acid transport. Neurology 81: 681-687. 23873973
Mojzita, D., and S. Hohmann. (2006). Pdc2 coordinates expression of the THI regulon in the yeast Saccharomyces cerevisiae. Mol. Genet. Genomics. 276: 147-161. 16850348
Mordoch, S.S., D. Granot, M. Lebendiker, and S. Schuldiner. (1999). Scanning cysteine accessibility of EmrE, an H+-coupled multidrug transporter from Escherichia coli, reveals a hydrophobic pathway for solutes. J. Biol. Chem. 274: 19480-19486. 10383465
Moskovskich, A., U. Goldmann, F. Kartnig, S. Lindinger, J. Konecka, G. Fiume, E. Girardi, and G. Superti-Furga. (2019). The transporters SLC35A1 and SLC30A1 play opposite roles in cell survival upon VSV virus infection. Sci Rep 9: 10471. 31320712
Müller, B., A. Fastner, J. Karmann, V. Mansch, T. Hoffmann, W. Schwab, M. Suter-Grotemeyer, D. Rentsch, E. Truernit, F. Ladwig, A. Bleckmann, T. Dresselhaus, and U.Z. Hammes. (2015). Amino Acid Export in Developing Arabidopsis Seeds Depends on UmamiT Facilitators. Curr. Biol. 25: 3126-3131. 26628011
Mullineaux, C.W., V. Mariscal, A. Nenninger, H. Khanum, A. Herrero, E. Flores, and D.G. Adams. (2008). Mechanism of intercellular molecular exchange in heterocyst-forming cyanobacteria. EMBO. J. 27: 1299-1308. 18388860
Muraoka, M., M. Kawakita, and N. Ishida. (2001). Molecular characterization of human UDP-glucuronic acid/UDP-N-acetylgalactosamine transporter, a novel nucleotide sugar transporter with dual substrate specificity. FEBS Lett. 495: 87-93. 11322953
Nasie I., Steiner-Mordoch S., Gold A. and Schuldiner S. (2010). Topologically random insertion of EmrE supports a pathway for evolution of inverted repeats in ion-coupled transporters. J Biol Chem. 285(20):15234-44. 20308069
Nessler, S., O. Friedrich, N. Bakouh, R.H. Fink, C.P. Sanchez, G. Planelles, and M. Lanzer. (2004). Evidence for activation of endogenous transporters in Xenopus laevis oocytes expressing the Plasmodium falciparum chloroquine resistance transporter, PfCRT. J. Biol. Chem. 279: 39438-39446. 15258157
Nicolet, Y. and C.L. Drennan. (2004). AdoMet radical proteins--from structure to evolution--alignment of divergent protein sequences reveals strong secondary structure element conservation. Nucleic Acids Res 32: 4015-4025. 15289575
Nies, D.H. (2003). Efflux-mediated heavy metal resistance in prokaryotes. FEMS Microbiol. Rev. 27: 313-339. 12829273
Ninio, S. and S. Schuldiner. (2003). Characterization of an archaeal multidrug transporter with a unique amino acid composition. J. Biol. Chem. 278: 12000-12005. 12551892
Nishikawa, A., J.B. Poster, Y. Jigami, and N. Dean. (2002). Molecular and phenotypic analysis of CaVRG4, encoding an essential Golgi apparatus GDP-mannose transporter. J. Bacteriol. 184: 29-42. 11741841
Nishimura, M., S. Suzuki, T. Satoh, and S. Naito. (2009). Tissue-specific mRNA expression profiles of human solute carrier 35 transporters. Drug Metab Pharmacokinet 24: 91-99. 19252338
Nishino, K. and A. Yamaguchi. (2001). Analysis of a complete library of putative drug transporter genes in Escherichia coli. J. Bacteriol. 183: 5803-5812. 11566977
Norambuena, L., L. Marchant, P. Berninsone, C.B. Hirschberg, H. Silva, and A. Orellana. (2002). Transport of UDP-galactose in plants. Identification and functional characterization of AtUTr1, an Arabidopsis thaliana UDP-galactos/UDP-glucose transporter. J. Biol. Chem. 277: 32923-32929. 12042319
Ondo, K., H. Arakawa, M. Nakano, T. Fukami, and M. Nakajima. (2020). SLC35B1 significantly contributes to the uptake of UDPGA into the endoplasmic reticulum for glucuronidation catalyzed by UDP-glucuronosyltransferases. Biochem Pharmacol 175: 113916. 32179043
Ong, Y.S., A. Lakatos, J. Becker-Baldus, K.M. Pos, and C. Glaubitz. (2013). Detecting substrates bound to the secondary multidrug efflux pump EmrE by DNP-enhanced solid-state NMR. J. Am. Chem. Soc. 135: 15754-15762. 24047229
Paguio, M.F., M. Cabrera, and P.D. Roepe. (2009). Chloroquine transport in Plasmodium falciparum. 2. Analysis of PfCRT-mediated drug transport using proteoliposomes and a fluorescent chloroquine probe. Biochemistry 48: 9482-9491. 19725576
Pan, Q., B. Cui, F. Deng, J. Quan, G.J. Loake, and W. Shan. (2016). RTP1 encodes a novel endoplasmic reticulum (ER)-localized protein in Arabidopsis and negatively regulates resistance against biotrophic pathogens. New Phytol 209: 1641-1654. 26484750
Park, J., B. Tefsen, M.J. Heemskerk, E.L. Lagendijk, C.A. van den Hondel, I. van Die, and A.F. Ram. (2015). Identification and functional analysis of two Golgi-localized UDP-galactofuranose transporters with overlapping functions in Aspergillus niger. BMC Microbiol 15: 253. 26526354
Paulsen, I.T., R.A. Skurray, R. Tam, M.H. Saier, Jr., R.J. Turner, J.H. Weiner, E.B. Goldberg, and L.L. Grinius. (1996). The SMR family: a novel family of multidrug efflux proteins involved with the efflux of lipophilic drugs. Mol. Microbiol. 19: 1167-1175. 8730859
Pimpat, Y., N. Saralamba, U. Boonyuen, S. Pukrittayakamee, F. Nosten, F. Smithuis, N.P.J. Day, A.M. Dondorp, and M. Imwong. (2020). Genetic analysis of the orthologous crt and mdr1 genes in Plasmodium malariae from Thailand and Myanmar. Malar J 19: 315. 32867773
Poulsen, B.E. and C.M. Deber. (2012). Drug efflux by a small multidrug resistance protein is inhibited by a transmembrane peptide. Antimicrob. Agents Chemother. 56: 3911-3916. 22526304
Poulsen, B.E., A. Rath, and C.M. Deber. (2009). The assembly motif of a bacterial small multidrug resistance protein. J. Biol. Chem. 284: 9870-9875. 19224913
Prentice, L.M., X. d''Anglemont de Tassigny, S. McKinney, T. Ruiz de Algara, D. Yap, G. Turashvili, S. Poon, M. Sutcliffe, P. Allard, A. Burleigh, J. Fee, D.G. Huntsman, W.H. Colledge, and S.A. Aparicio. (2011). The testosterone-dependent and independent transcriptional networks in the hypothalamus of Gpr54 and Kiss1 knockout male mice are not fully equivalent. BMC Genomics 12: 209. 21527035
Qazi, S.J.S. and R.J. Turner. (2018). Influence of quaternary cation compound on the size of thesmall multidrug resistance protein, EmrE. Biochem Biophys Rep 13: 129-140. 29552647
Qi, W., X.X. Li, Y.H. Guo, Y.Z. Bao, N. Wang, X.G. Luo, C.D. Yu, and T.C. Zhang. (2020). Integrated metabonomic-proteomic analysis reveals the effect of glucose stress on metabolic adaptation of Lactococcus lactis ssp. lactis CICC23200. J Dairy Sci. [Epub: Ahead of Print] 32684472
Radi, M.S., J.E. SalcedoSora, S.H. Kim, S. Sudarsan, A.V. Sastry, D.B. Kell, M.J. Herrgård, and A.M. Feist. (2022). Membrane transporter identification and modulation via adaptive laboratory evolution. Metab Eng 72: 376-390. 35598887
Rautengarten, C., B. Ebert, I. Moreno, H. Temple, T. Herter, B. Link, D. Doñas-Cofré, A. Moreno, S. Saéz-Aguayo, F. Blanco, J.C. Mortimer, A. Schultink, W.D. Reiter, P. Dupree, M. Pauly, J.L. Heazlewood, H.V. Scheller, and A. Orellana. (2014). The Golgi localized bifunctional UDP-rhamnose/UDP-galactose transporter family of Arabidopsis. Proc. Natl. Acad. Sci. USA 111: 11563-11568. 25053812
Reed, J.A., P.A. Wilkinson, H. Patel, M.A. Simpson, A. Chatonnet, D. Robay, M.A. Patton, A.H. Crosby, and T.T. Warner. (2005). A novel NIPA1 mutation associated with a pure form of autosomal dominant hereditary spastic paraplegia. Neurogenetics 6: 79-84. 15711826
Reverchon, S., W. Nasser, and J. Robert-Baudouy. (1994). pecS: a locus controlling pectinase, cellulase and blue pigment production in Erwinia chrysanthemi. Mol. Microbiol. 11: 1127-1139. 8022282
Rodionova, I.A., F. Heidari Tajabadi, Z. Zhang, D.A. Rodionov, and M.H. Saier, Jr. (2019). A Riboflavin Transporter in Bdellovibrio exovorous JSS. J. Mol. Microbiol. Biotechnol. 29: 27-34. 31509826
Rodrigues, L., C. Villellas, R. Bailo, M. Viveiros, and J.A. Aínsa. (2013). Role of the Mmr efflux pump in drug resistance in Mycobacterium tuberculosis. Antimicrob. Agents Chemother. 57: 751-757. 23165464
Rodriguez, P.Q., A. Oddsson, L. Ebarasi, B. He, K. Hultenby, A. Wernerson, C. Betsholtz, K. Tryggvason, and J. Patrakka. (2015). Knockdown of Tmem234 in zebrafish results in proteinuria. Am. J. Physiol. Renal Physiol 309: F955-966. 26377798
Rong, C., R. Zhang, Y. Liu, Z. Chang, Z. Liu, Y. Ding, and C. Ding. (2024). Purine permease (PUP) family gene PUP11 positively regulates the rice seed setting rate by influencing seed development. Plant Cell Rep 43: 112. 38568250
Rouanet, C. and W. Nasser. (2001). The PecM protein of the phytopathogenic bacterium Erwinia chrysanthemi, membrane topology and possible involvement in the efflux of the blue pigment indigoidine. J. Mol. Microbiol. Biotechnol. 3: 309-318. 11321588
Roy, S.K., Y. Chiba, M. Takeuchi, and Y. Jigami. (2000). Characterization of Yeast Yea4p, a uridine diphosphate-N-acetylglucosamine transporter localized in the endoplasmic reticulum and required for chitin synthesis. J. Biol. Chem. 275: 13580-13587. 10788474
Sá, J.M., M.M. Yamamoto, C. Fernandez-Becerra, M.F. de Azevedo, J. Papakrivos, B. Naudé, T.E. Wellems, and H.A. Del Portillo. (2006). Expression and function of pvcrt-o, a Plasmodium vivax ortholog of pfcrt, in Plasmodium falciparum and Dictyostelium discoideum. Mol Biochem Parasitol 150: 219-228. 16987557
Saleh, M., D.C. Bay, and R.J. Turner. (2018). Few conserved amino acids in the small multidrug resistance transporter EmrE influence drug polyselectivity. Antimicrob. Agents Chemother. [Epub: Ahead of Print] 29866867
Santiviago, C.A., J.A. Fuentes, S.M. Bueno, A.N. Trombert, A.A. Hildago, L.T. Socias, P. Youderian, and G.C. Mora. (2002). The Salmonella enterica sv. Typhimurium smvA, yddG,and ompD (porin) genes are required for the efficient efflux of methyl viologen. Mol. Microbiol. 46: 687-698. 12410826
Sarkar, S.K., A. Bhattacharyya, and S.M. Mandal. (2015). YnfA , a SMR family efflux pump is abundant in Escherichia coli isolates from urinary infection. Indian J. Med. Microbiol. 33: 139-142. 25560019
Schäffers, O.J.M., J.G.J. Hoenderop, R.J.M. Bindels, and J.H.F. de Baaij. (2018). The rise and fall of novel renal magnesium transporters. Am. J. Physiol. Renal Physiol 314: F1027-F1033. 29412701
Schmidt, A., N. Baumann, A. Schwarzkopf, W.B. Frommer, and M. Desimone. (2006). Comparative studies on Ureide Permeases in Arabidopsis thaliana and analysis of two alternative splice variants of AtUPS5. Planta 224: 1329-1340. 16738859
Schmidt, A., Y.H. Su, R. Kunze, S. Warner, M. Hewitt, R.D. Slocum, U. Ludewig, W.B. Frommer, and M. Desimone. (2004). UPS1 and UPS2 from Arabidopsis mediate high affinity transport of uracil and 5-fluorouracil. J. Biol. Chem. 279: 44817-44824. 15308648
Schwaiger, M., M. Lebendiker, H. Yerushalmi, M. Coles, A. Gröger, C. Schwarz, S. Schuldiner, and H. Kessler. (1998). NMR investigation of the multidrug transporter EmrE, an integral membrane protein. Eur J Biochem 254: 610-619. 9688273
Schwarz, M., A. Gross, T. Steinkamp, U.-I. Flügge, and R. Wagner. (1994). Ion channel properties of the reconstituted chloroplast triose phosphate/phosphate translocator. J. Biol. Chem. 269: 29481-29489. 7525584
Schwarzbaum, P.J., J. Schachter, and L.M. Bredeston. (2022). The broad range di- and trinucleotide exchanger SLC35B1 displays asymmetrical affinities for ATP transport across the ER membrane. J. Biol. Chem. 101537. [Epub: Ahead of Print] 35041824
Seino, J., K. Ishii, T. Nakano, N. Ishida, M. Tsujimoto, Y. Hashimoto, and S. Takashima. (2010). Characterization of rice nucleotide sugar transporters capable of transporting UDP-galactose and UDP-glucose. J Biochem 148: 35-46. 20305274
Selva, E.M., K. Hong, G.H. Baeg, S.M. Beverley, S.J. Turco, N. Perrimon, and U. Häcker. (2001). Dual role of the fringe connection gene in both heparan sulphate and fringe-dependent signalling events. Nat. Cell Biol. 3: 809-815. 11533660
Sesma, J.I., C.R. Esther, Jr, S.M. Kreda, L. Jones, W. O'Neal, S. Nishihara, R.A. Nicholas, and E.R. Lazarowski. (2009). Endoplasmic reticulum/golgi nucleotide sugar transporters contribute to the cellular release of UDP-sugar signaling molecules. J. Biol. Chem. 284: 12572-12583. 19276090
Seurig, M., M. Ek, G. von Heijne, and N. Fluman. (2019). Dynamic membrane topology in an unassembled membrane protein. Nat Chem Biol 15: 945-948. 31501590
Shi, X., Y. Li, P. Yan, Y. Shi, and J. Lai. (2020). Weighted gene co-expression network analysis to explore the mechanism of heroin addiction in human nucleus accumbens. J. Cell. Biochem. 121: 1870-1879. 31692007
Short, B. (2024). SMR transporters meet the challenge of metformin metabolites. J Gen Physiol 156:. 38324209
Snyder, N.A., A. Kim, L. Kester, A.N. Gale, C. Studer, D. Hoepfner, S. Roggo, S.B. Helliwell, and K.W. Cunningham. (2019). Auxin-Inducible Depletion of the Essentialome Suggests Inhibition of TORC1 by Auxins and Inhibition of Vrg4 by SDZ 90-215, a Natural Antifungal Cyclopeptide. G3 (Bethesda) 9: 829-840. 30670608
Song, Z. (2013). Roles of the nucleotide sugar transporters (SLC35 family) in health and disease. Mol Aspects Med 34: 590-600. 23506892
Sosicka, P., D. Maszczak-Seneczko, B. Bazan, Y. Shauchuk, B. Kaczmarek, and M. Olczak. (2017). An insight into the orphan nucleotide sugar transporter SLC35A4. Biochim. Biophys. Acta. [Epub: Ahead of Print] 28167211
Spreacker, P.J., N.E. Thomas, W.F. Beeninga, M. Brousseau, C.J. Porter, K.M. Hibbs, and K.A. Henzler-Wildman. (2022). Activating alternative transport modes in a multidrug resistance efflux pump to confer chemical susceptibility. Nat Commun 13: 7655. 36496486
Srinivasan, V.B., G. Rajamohan, and W.A. Gebreyes. (2009). Role of AbeS, a novel efflux pump of the SMR family of transporters, in resistance to antimicrobial agents in Acinetobacter baumannii. Antimicrob. Agents Chemother. 53: 5312-5316. 19770280
Suda, T., S. Kamiyama, M. Suzuki, N. Kikuchi, K. Nakayama, H. Narimatsu, Y. Jigami, T. Aoki, and S. Nishihara. (2004). Molecular cloning and characterization of a human multisubstrate specific nucleotide-sugar transporter homologous to Drosophila fringe connection. J. Biol. Chem. 279: 26469-26474. 15082721
Swierkowska, J., J.A. Karolak, T. Gambin, M. Rydzanicz, A. Frajdenberg, M. Mrugacz, M. Podfigurna-Musielak, P. Stankiewicz, J.R. Lupski, and M. Gajecka. (2021). Variants in FLRT3 and SLC35E2B identified using exome sequencing in seven high myopia families from Central Europe. Adv Med Sci 66: 192-198. 33711669
Takeshima-Futagami T., Sakaguchi M., Uehara E., Aoki K., Ishida N., Sanai Y., Sugahara Y. and Kawakita M. (2012). Amino acid residues important for CMP-sialic acid recognition by the CMP-sialic acid transporter: analysis of the substrate specificity of UDP-galactose/CMP-sialic acid transporter chimeras. Glycobiology. 22(12):1731-40. 22833315
Takita, Y., Y. Ohya, and Y. Anraku. (1995). The CLS2 gene encodes a protein with multiple membrane-spanning domains that is important Ca2+ tolerance in yeast. Mol. Gen. Genet. 246: 269-281. 7854312
Tan, L.L., T.Y. Lau, W. Timothy, and D. Prabakaran. (2014). Full-length sequence analysis of chloroquine resistance transporter gene in Plasmodium falciparum isolates from Sabah, Malaysia. ScientificWorldJournal 2014: 935846. 25574497
Tanida, I., Y. Takita, A. Hasegawa, Y. Ohya, and Y. Anraku. (1996). Yeast Cls2p/Csg2p localized on the endoplasmic reticulum membrane regulates a non-exchangeable intracellular Ca2+ pool cooperatively with calcineurin. FEBS Lett. 379: 38-42. 8566225
Tate, C.G., E.R.S. Kunji, M. Lebendiker, and S. Schuldiner. (2001). The projection structure of EmrE, a proton-linked multidrug transporter from Escherichia coli, at 7 Å resolution. EMBO J. 20: 77-81. 11226157
Tate, C.G. and P.J. Henderson. (1993). Membrane topology of the L-rhamnose-H+ transport protein (RhaT) from enterobacteria. J. Biol. Chem. 268: 26850-26857. 8262918
Tate, C.G., J.A. Muiry, and P.J. Henderson. (1992). Mapping, cloning, expression, and sequencing of the rhaT gene, which encodes a novel L-rhamnose-H+ transport protein in Salmonella typhimurium and Escherichia coli. J. Biol. Chem. 267: 6923-6932. 1551902
Thongchuang, M., P. Pongsawasdi, Y. Chisti, and K. Packdibamrung. (2012). Design of a recombinant Escherichia coli for producing L-phenylalanine from glycerol. World J Microbiol Biotechnol 28: 2937-2943. 22806734
Toscanini, M.A., M.B. Favarolo, F.L. Gonzalez Flecha, B. Ebert, C. Rautengarten, and L.M. Bredeston. (2019). Conserved Glu-47 and Lys-50 residues are critical for UDP-N-acetylglucosamine/UMP antiport activity of the mouse Golgi-associated transporter SLC35A3. J. Biol. Chem. [Epub: Ahead of Print] 31118275
Tran, C.V. and M.H. Saier, Jr. (2004). The principal chloroquine resistance protein of Plasmodium falciparum is a member of the drug/metabolite transporter superfamily. Microbiology 150: 1-3. 14702390
Tsuchiya, H., S. Doki, M. Takemoto, T. Ikuta, T. Higuchi, K. Fukui, Y. Usuda, E. Tabuchi, S. Nagatoishi, K. Tsumoto, T. Nishizawa, K. Ito, N. Dohmae, R. Ishitani, and O. Nureki. (2016). Structural basis for amino acid export by DMT superfamily transporter YddG. Nature 534: 417-420. 27281193
Tucker, A.M., H.H. Winkler, L.O. Driskell, and D.O. Wood. (2003). S-Adenosylmethionine transport in Rickettsia prowazekii. J. Bacteriol. 185: 3031-3035. 0
Ubarretxena-Belandia, I., J.M. Baldwin, S. Schuldiner, and C.G. Tate. (2003). Three-dimensional structure of the bacterial multidrug transporter EmrE shows it is an asymmetric homodimer. EMBO J. 22: 6175-6181. 14633977
Uemura, S., A. Kihara, S. Iwaki, J. Inokuchi, and Y. Igarashi. (2007). Regulation of the transport and protein levels of the inositol phosphorylceramide mannosyltransferases Csg1 and Csh1 by the Ca2+-binding protein Csg2. J. Biol. Chem. 282: 8613-8621. 17220303
Västermark, &.#.1.9.7.;., M.S. Almén, M.W. Simmen, R. Fredriksson, and H.B. Schiöth. (2011). Functional specialization in nucleotide sugar transporters occurred through differentiation of the gene cluster EamA (DUF6) before the radiation of Viridiplantae. BMC Evol Biol 11: 123. 21569384
Verstraeten, N., W.J. Knapen, C.I. Kint, V. Liebens, B. Van den Bergh, L. Dewachter, J.E. Michiels, Q. Fu, C.C. David, A.C. Fierro, K. Marchal, J. Beirlant, W. Versées, J. Hofkens, M. Jansen, M. Fauvart, and J. Michiels. (2015). Obg and Membrane Depolarization Are Part of a Microbial Bet-Hedging Strategy that Leads to Antibiotic Tolerance. Mol. Cell 59: 9-21. 26051177
Vía, P., J. Badía, L. Baldomà, N. Obradors, and J. Aguilar. (1996). Transcriptional regulation of the Escherichia coli rhaT gene. Microbiology (Reading) 142(Pt7): 1833-1840. 8757746
Vitreschak, A.G., D.A. Rodionov, A.A. Mironov, and M.S. Gelfand. (2002). Regulation of riboflavin biosynthesis and transport genes in bacteria by transcriptional and translational attenuation. Nucleic Acids Res 30: 3141-3151. 12136096
Wang, J., L.K. Cheng, J. Wang, Q. Liu, T. Shen, and N. Chen. (2013). Genetic engineering of Escherichia coli to enhance production of L-tryptophan. Appl. Microbiol. Biotechnol. 97: 7587-7596. 23775271
Wang, N., Y. Zhou, Z. Zuo, R. Wang, J. Li, T. Han, and B. Yang. (2021). Construction of a competing endogenous RNA network related to the prognosis of cholangiocarcinoma and comprehensive analysis of the immunological correlation. J Gastrointest Oncol 12: 2287-2309. 34790393
Wang, Z.A., C.L. Griffith, M.L. Skowyra, N. Salinas, M. Williams, E.J. Maier, S.R. Gish, H. Liu, M.R. Brent, and T.L. Doering. (2014). Cryptococcus neoformans dual GDP-mannose transporters and their role in biology and virulence. Eukaryot. Cell. 13: 832-842. 24747214
Wei, Z.B., Y.F. Yuan, F. Jaouen, M.S. Ma, C.J. Hao, Z. Zhang, Q. Chen, Z. Yuan, L. Yu, C. Beurrier, and W. Li. (2016). SLC35D3 increases autophagic activity in midbrain dopaminergic neurons by enhancing BECN1-ATG14-PIK3C3 complex formation. Autophagy 12: 1168-1179. 27171858
Wellems, T.E. (2002). Plasmodium chloroquine resistance and the search for a replacement antimalarial drug. Science 298: 124-126. 12364789
Williams, L.E. and A.J. Miller. (2001). Transporters responsible for the uptake and partitioning of nitrogenous solutes. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52: 659-688. 11337412
Wiser, M.F. (2024). The Digestive Vacuole of the Malaria Parasite: A Specialized Lysosome. Pathogens 13:. 38535526
Woodall, N.B., S. Hadley, Y. Yin, and J.U. Bowie. (2017). Complete topology inversion can be part of normal membrane protein biogenesis. Protein. Sci. [Epub: Ahead of Print] 28168866
Woodall, N.B., Y. Yin, and J.U. Bowie. (2015). Dual-topology insertion of a dual-topology membrane protein. Nat Commun 6: 8099. 26306475
Wunderlich, J. (2022). Updated List of Transport Proteins in. Front Cell Infect Microbiol 12: 926541. 35811673
Xie, M., F. Wang, B. Chen, Z. Wu, C. Chen, and J. Xu. (2023). Systematic pan-cancer analysis identifies SLC35C1 as an immunological and prognostic biomarker. Sci Rep 13: 5331. 37005450
Xu, D., N.C.H. Sanden, L.L. Hansen, Z.M. Belew, S.R. Madsen, L. Meyer, M.E. Jørgensen, P. Hunziker, D. Veres, C. Crocoll, A. Schulz, H.H. Nour-Eldin, and B.A. Halkier. (2023). Export of defensive glucosinolates is key for their accumulation in seeds. Nature. [Epub: Ahead of Print] 37076627
Yamada, Y., J. Sakuma, I. Takeuchi, Y. Yasukochi, K. Kato, M. Oguri, T. Fujimaki, H. Horibe, M. Muramatsu, M. Sawabe, Y. Fujiwara, Y. Taniguchi, S. Obuchi, H. Kawai, S. Shinkai, S. Mori, T. Arai, and M. Tanaka. (2017). Identification of TNFSF13, SPATC1L, SLC22A25 and SALL4 as novel susceptibility loci for atrial fibrillation by an exome‑wide association study. Mol Med Rep 16: 5823-5832. 28849223
Yamamoto, K., G. Nonaka, T. Ozawa, K. Takumi, and A. Ishihama. (2015). Induction of the Escherichia coli yijE gene expression by cystine. Biosci. Biotechnol. Biochem. 79: 218-222. 25346166
Yan, A., Z. Guan, and C.R. Raetz. (2007). An undecaprenyl phosphate-aminoarabinose flippase required for polymyxin resistance in Escherichia coli. J. Biol. Chem. 282: 36077-36089. 17928292
Yan, Q., E. Forno, A. Cardenas, C. Qi, Y.Y. Han, E. Acosta-Pérez, S. Kim, R. Zhang, N. Boutaoui, G. Canino, J.M. Vonk, C.J. Xu, W. Chen, E. Oken, D.R. Gold, G.H. Koppelman, and J.C. Celedón. (2020). Exposure to violence, chronic stress, nasal DNA methylation, and atopic asthma in children. medRxiv. 33173928
Yerushalmi, H. and S. Schuldiner. (2000). A model for coupling of H+ and substrate fluxes based on "time-sharing" of a common binding site. Amer. Chem. Soc. 39: 14711-14719. 0
Zakataeva, N.P., E.A. Kutukova, S.V. Gronskiĭ, P.V. Troshin, V.A. Livshits, and V.V. Aleshin. (2006). [Export of metabolites by the proteins of the DMT and RhtB families and its possible role in intercellular communication]. Mikrobiologiia 75: 509-520. 17025177
Zang, X.L., W.Q. Han, F.P. Yang, K.D. Ji, J.G. Wang, P.J. Gao, G. He, and S.N. Wu. (2016). Association of a SNP in SLC35F3 Gene with the Risk of Hypertension in a Chinese Han Population. Front Genet 7: 108. 27379158
Zhang, K., M.J. Huentelman, F. Rao, E.I. Sun, J.J. Corneveaux, A.J. Schork, Z. Wei, J. Waalen, J.P. Miramontes-Gonzalez, C.M. Hightower, A.X. Maihofer, M. Mahata, T. Pastinen, G.B. Ehret, , N.J. Schork, E. Eskin, C.M. Nievergelt, M.H. Saier, Jr, and D.T. O''Connor. (2014). Genetic implication of a novel thiamine transporter in human hypertension. J Am Coll Cardiol 63: 1542-1555. 24509276
Zhang, P., R. Haryadi, K.F. Chan, G. Teo, J. Goh, N.A. Pereira, H. Feng, and Z. Song. (2012). Identification of functional elements of the GDP-fucose transporter SLC35C1 using a novel Chinese hamster ovary mutant. Glycobiology 22: 897-911. 22492235
Zhang, S., Q. Li, H. Yuan, L. Ren, X. Liang, S. Li, S. Lv, and H. Jiang. (2022). Solute Carrier Family 35 Member F2 Regulates Cisplatin Resistance and Promotes Malignant Progression of Pancreatic Cancer by Regulating RNA Binding Motif Protein 14. J Oncol 2022: 5091154. 35669242
Zhang, Z., C. Ma, O. Pornillos, X. Xiu, G. Chang, and M.H. Saier, Jr. (2007). Functional characterization of the heterooligomeric EbrAB multidrug efflux transporter of Bacillus subtilis. Biochemistry 46: 5218-5225. 17417881
Zhou, H., Q. Shen, J. Fu, F. Jiang, L. Wang, and Y. Wang. (2020). Analysis of lncRNA UCA1-related downstream pathways and molecules of cisplatin resistance in lung adenocarcinoma. J Clin Lab Anal 34: e23312. 32249461