Electrolyte Blood Press. 2021 Jun;19(1):1-9. English.
Published online Jun 30, 2021.
Copyright © 2021 Korean Society for Electrolyte and Blood Pressure Research
Review

Urate Transporters in the Kidney: What Clinicians Need to Know

Sungjin Chung,1 and Gheun-Ho Kim2
    • 1Department of Internal Medicine, College of Medicine, The Catholic University of Korea, Seoul, Republic of Korea.
    • 2Department of Internal Medicine, Hanyang University College of Medicine, Seoul, Republic of Korea.
Received May 03, 2021; Accepted May 27, 2021.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/4.0/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Abstract

Urate is produced in the liver by the degradation of purines from the diet and nucleotide turnover and excreted by the kidney and gut. The kidney is the major route of urate removal and has a pivotal role in the regulation of urate homeostasis. Approximately 10% of the glomerular filtered urate is excreted in the urine, and the remainder is reabsorbed by the proximal tubule. However, the transport of urate in the proximal tubule is bidirectional: reabsorption and secretion. Thus, an increase in reabsorption or a decrease in secretion may induce hyperuricemia. In contrast, a decrease in reabsorption or an increase in secretion may result in hyperuricosuria. In the proximal tubule, urate reabsorption is mainly mediated by apical URAT1 (SLC22A12) and basolateral GLUT9 (SLC2A9) transporter. OAT4 (SLC22A11) also acts in urate reabsorption in the apical membrane, and its polymorphism is associated with the risk of hyperuricemia. Renal hypouricemia is caused by SLC22A12 or SLC2A9 loss-of-function mutations, and it may be complicated by exercise-induced acute kidney injury. URAT1 and GLUT9 are also drug targets for uricosuric agents. Sodium-glucose cotransporter inhibitors may induce hyperuricosuria by inhibiting GLUT9b located in the apical plasma membrane. Urate secretion is mediated by basolateral OAT1 (SLC22A6) and OAT3 (SLC22A8) and apical ATP-binding cassette super-family G member 2 (ABCG2), NPT1 (SLC17A1), and NPT4 (SLC17A3) transporter in the proximal tubule. NPT1 and NPT4 may be key players in renal urate secretion in humans, and deletion of SLC22A6 and SLC22A8 in mice leads to decreased urate excretion. Dysfunctional variants of ABCG2 inhibit urate secretion from the gut and kidney and may cause gout. In summary, the net result of urate transport in the proximal tubule is determined by the dominance of transporters between reabsorption (URAT1, OAT4, and GLUT9) and secretion (ABCG2, NPT1, NPT4, OAT1, and OAT3).

Keywords
Gout; Hyperuricosuria; Hypouricemia; Proximal tubule; Uric acid transport

Urate Homeostasis

Uric acid or urate homeostasis in humans is determined by the balance between hepatic production and renal and intestinal excretion1). Purines are metabolized to urate in the liver, and their sources are both exogenous (dietary intake) and endogenous (cellular metabolism of nucleic acids). The enzyme xanthine oxidase (XO) acts on the final step of purine metabolism from xanthine to urate, and XO inhibitors such as allopurinol and febuxostat are clinically useful as urate-lowering therapy. In most mammals, urate is converted by uricase or urate oxidase into allantoin, which is water-soluble and easily excreted in urine. In humans and apes, however, urate is the end product of purine metabolism because of nonsense mutations in uricase, and plasma urate levels are three to ten times higher than in other mammals2).

Uric acid has two dissociable protons with an aqueous pKa1 of 5.4 and pKa2 of 9.8 (Fig. 1). Consequently, at the physiological pH of 7.4, uric acid in the extracellular fluid is predominantly (98-99%) found as a deprotonated urate anion3). The divalent urate anion is practically nonexistent in the body because of the very high pKa2, and thus the term urate is generally used to refer to monovalent urate in the biomedical literature. Because the ratio of urate-to-uric acid in the circulation remains constant with constant pH, the terms urate and uric acid are often used interchangeably to refer to the total pool of uric acid, dissociated and undissociated1).

Fig. 1
Uric acid pKa and formation of urate salts.
At physiological pH, uric acid is predominantly found as the deprotonated urate anion. Adapted from Reference 3.

Abnormal urate homeostasis can be classified into hyperuricemia and hypouricemia. Hyperuricemia may be induced by increased urate production in the liver or decreased urate excretion through the kidney and/or gut. Although a purine-rich diet may increase hepatic production of uric acid, diet alone is generally insufficient to cause hyperuricemia. Lesch-Nyhan syndrome is caused by hypoxanthine-guanine phosphoribosyltransferase (HPRT) deficiency and resultant uric acid overproduction and is presented with dystonia, gout, intellectual disability, and self-mutilation4). HGPRT is an enzyme catalyzing the conversion of hypoxanthine to inosine monophosphate and guanine to guanosine monophosphate, generating purine nucleotides through the purine salvage pathway5). Urate underexcretion through the gut and/or kidney is the other important etiology of hyperuricemia, and intestinal and renal urate transporters may be involved in these pathological conditions. Diet and genetic polymorphisms in the renal transporters of urate seem to be the main causal factors of primary gout6).

Theoretically, hypouricemia can be caused by decreased urate production in the liver or increased urate excretion through the kidney and/or gut. Malnutrition may be associated with hypouricemia because of low dietary purine intake. Molecular defects in critical pathways involving urate synthesis in the liver or urate reabsorption in the kidney may cause hypouricemia. Medications affecting urate production in the liver (e.g., allopurinol, febuxostat) and urate transport through the gut and kidney (e.g., probenecid, benzbromarone) may lower serum uric acid levels.

Urate Transporters in the Kidney

The kidney is the major route of urate removal, typically responsible for 70-80% of daily urate excretion, and the remainder is eliminated by intestinal secretion7). Plasma urate is freely filtered at the glomeruli, and fractional excretion of urate is approximately 10% in adult humans1). Thus, 90% of the glomerular filtrates are normally reabsorbed along the nephron, but this is the net result of reabsorption and secretion in the proximal tubule.

Figure 2 illustrates major urate transporters in the proximal tubule. Considering the usual ranges of fractional excretion of urate, transporters mediating urate reabsorption may be more influential than those mediating urate secretion. The former includes apically located urate transporter 1 (URAT1), organic anion transporter 4 (OAT4), and OAT10, and basolaterally located glucose transporter 9 (GLUT9). URAT1, encoded by the SLC22A12 gene, is the primary urate/anion exchanger responsible for luminal urate reabsorption. For the exchange of urate with organic anions, it is coupled with sodium monocarboxylate cotransporter 1 (SMCT1) and SMCT2 at the apical membrane of the proximal tubule. With URAT1, organic anions such as nicotinate and pyrazinoate have a higher affinity than other anions such as lactate, β-hydroxybutyrate, acetoacetate, chloride, and nitrate8).

Fig. 2
Major urate transporters for bidirectional pathways in the proximal tubule.
(A) Na+-dependent anion transport by SMCT1 and SMCT2 increases intracellular concentrations of anions that exchange with luminal urate by URAT1. GLUT9 acts as the basolateral exit for urate reabsorption. (B) OAT1 and OAT3 transport urate through the basolateral membrane in exchange with α-ketoglutarate (α-KG). At the apical membrane, urate is secreted via ABCG2, NPT1, and/or NPT4.

OAT4 and OAT10 may act as urate/anion exchangers like URAT1. OAT10, encoded by SLC22A13, is expressed at the messenger RNA (mRNA) level in the human kidney, but not at the protein level9). Its function as a urate/monocarboxylate exchanger was demonstrated only in vitro with a lower affinity for urate than URAT110). Human OAT10 is also expressed to a weak extent in the intestine but its function remains unclear11). Unlike URAT1 and OAT10, OAT4, encoded by SLC22A11, mediates urate/dicarboxylate exchange, with a lower estimated affinity for urate compared with URAT112). Thus, URAT1 appears to play the most important role in apical entrance for urate reabsorption.

Interestingly, URAT1 was also localized to the cell membrane of human vascular smooth muscle cells. This finding is compatible with the role of uric acid in cardiovascular disease, suggesting that uric acid enters the human vascular smooth muscle cell via URAT113).

GLUT9, encoded by the SLC2A9 gene, functions as the major mechanism for the basolateral exit of urate, allowing for its reabsorption into the blood10). Human GLUT9 has two splice variants with distinct N-terminal isoforms; GLUT9a is expressed in multiple tissues including the small intestine and colon, while GLUT9b (also termed GLUT9ΔN) is highly expressed in the kidney, and to a lesser extent in the liver14). Their unique N-termini are responsible for differential localization, with GLUT9a localizing to the basolateral membrane of the proximal tubule and GLUT9b located in the apical membrane of the proximal tubule14) or collecting duct15). In the murine kidney, Glut9a is expressed weakly in the proximal tubule whereas Glut9b is present in the distal convoluted tubule16). Physiologic action and species-specific intrarenal localization of GLUT9b remain to be confirmed.

Another set of transporters mediate urate secretion in the proximal tubule. ATP-binding cassette super-family G member 2 (ABCG2), also known as the breast cancer-resistant protein (BCRP), was initially identified as a transporter of drugs but has a role in the ATP-dependent urate efflux17). It is located in the apical membrane of the proximal tubule for secretion of urate, along with other ABC transporters such as multidrug resistance protein 2 (MRP2) and MRP418). ABCG2 is also highly expressed in intestinal tissue, where it excretes up to one-third of all uric acid and is thus thought to be the main extrarenal site of uric acid elimination19). To a lesser extent, MRP2 also called ATP-binding cassette sub-family C member 2 (ABCC2) and MRP4 also called ATP-binding cassette sub-family C member 4 (ABCC4) are expressed in the intestine to efflux organic anions including urate11).

NPT1, encoded by SLC17A1, is a sodium-phosphate cotransporter located in the apical membrane of the proximal tubule in the human kidney and was identified as a candidate protein in mediating the electrogenic exit of urate20). The highly related NPT4, encoded by SLC17A3, is located at the apical membrane of the proximal tubule and mediates voltage-dependent urate transport21). It was also reported to exist in the intestines and liver22). On the other hand, NPT5, encoded by SLC17A4, is expressed in the intestinal mucosae but not in the kidney23). These findings may indicate a possible involvement of the NPT homologs in urate efflux from the kidney and intestinal tract.

OAT1 and OAT3, encoded by SLC22A6 and SLC22A8, respectively, are located in the basolateral membrane of the proximal tubule for urate uptake and overall function in renal urate secretion. They exchange urate and other organic anions for intracellular dicarboxylates such as α-ketoglutarate, coupled to intracellular sodium transport through the sodium-dicarboxylate cotransporter 3 (NaDC-3) encoded by SLC13A35). OAT2, encoded by SLC22A7, similarly is located in the basolateral membrane of the proximal tubule24), and its activity as a low-affinity urate transporter was found in transfected HEK293 cells25).

Renal Hypouricemia

Hypouricemia caused by a renal tubular defect was termed ‘renal hypouricemia’. The apical URAT1 and the basolateral GLUT9 are the main reabsorptive urate transporters, and loss-of-function mutations of the SLC22A12 and SLC2A9 gene can cause type 1 and type 2 renal hypouricemia, respectively26). Patients with renal hypouricemia can present with hematuria, urolithiasis, or exercise-induced acute kidney injury (AKI), and their fractional excretion of urate is much higher than 10% despite a very low level of serum urate. Among different mutations in the SLC22A12 gene, W258X (rs121907892) was predominant in patients with type 1 renal hypouricemia reported from both Japan27) and Korea28).

Whereas type 1 renal hypouricemia mainly occurs in Asian children, cases of type 2 renal hypouricemia were reported from various parts of the world, including Asia, Middle East, and Europe. Different mutations were clustered according to the regions, and patients with type 2 renal hypouricemia were often diagnosed during their adulthood29).

The reason why exercise-induced AKI can occur in patients with renal hypouricemia is unclear. One explanation could be that uric acid is a powerful antioxidant that scavenges singlet oxygen, oxygen radicals, and peroxynitrite7). The absence of uricase occurred approximately 30 million years ago in hominid evolution and was associated with the loss of the ability to synthesize ascorbic acid de novo30), suggesting that uric acid may have replaced ascorbic acid as an antioxidant31). Interestingly, uric acid may function as an antioxidant in plasma and may act as a pro-oxidant within the cell (in cardiovascular disease)32). In patients with hypouricemia, the antioxidant activity of uric acid is overwhelmed by the massive reactive oxygen species produced by strenuous exercise. Thus, loss of antioxidant activity may lead to vascular constriction and endothelial damages, progressing to AKI33).

Urate Transporters and the Risk of Hyperuricemia

1. URAT1

In addition to renal hypouricemia type 1, single nucleotide polymorphisms (SNPs) in the SLC22A12 gene may be associated with hyperuricemia or gout. A meta-analysis showed that the rs475688 polymorphism in the SLC22A12 gene was associated with gout susceptibility34). The rs559946 polymorphism was associated with increased hyperuricemia risk and might also contribute to gout development in Han Chinese men35). In the Vietnamese population, the rs11231825 polymorphism was associated with gout36). The rs3825017 polymorphism was associated with the risk of gout in a Czech population37). In the Korean Cancer Prevention Study-II cohort, the rs75786299, rs7929627, and rs3825017 SNPs were associated with hyperuricemia38).

2. GLUT9

In addition to renal hypouricemia type 2, the SLC2A9 gene may also be associated with hyperuricemia or gout. In particular, the nonsynonymous Arg265His (rs3733591) variant of SLC2A9 increases the risk for gout in some populations39). When a genome-wide association study (GWAS) was conducted in 6,881 Korean individuals, rs16890979 polymorphism was associated with hyperuricemia40). In Croatian, erman, and UK populations, three different SLC2A9 variants, rs1014290, rs6449213, and rs737267 were associated with gout41). Meta-analysis studies showed that the rs3733591 polymorphism in the SLC2A9 gene was associated with gout susceptibility42) whereas the rs12510549, rs16890979, and rs1014290 polymorphisms protect against the development of gout in Caucasians and/or Asians43).

Hyperuricemia is frequently accompanied by chronic kidney disease (CKD), but its causal relationship with CKD is unclear. Testa et al. reported that the rs734553 SNP in the SLC2A9 gene was not only associated with serum uric acid levels in 211 healthy individuals but also predicted the risk of CKD progression in a cohort of 755 patients, supporting the hypothesis that the link between uric acid and CKD progression is causal in nature44).

3. ABCG2

Dysfunction of ABCG2 causes hyperuricemia and raises the risk of gout because ABCG2 has a role in urate secretion in both proximal tubule and intestinal mucosa17). The intestinal excretion of urate is responsible for approximately 30% of total urate excretion45).

A common ABCG2 variant, Q141K (rs2231142), was associated with an increase of uric acid levels in multiple populations, including African Americans, Asians, Caucasians, and Pacific Islanders1). Sequencing of the ABCG2 gene in 90 hyperuricemia patients revealed several nonfunctional ABCG2 mutations, including Q126X46). When a GWAS was conducted with data from a Korean cohort (3,647 subjects recruited by the Korean National Institute of Health), rs2054576 in ABCG2 was associated with hyperuricemia47).

4. Other urate transporters

OAT4 (SLC22A11) is localized in the apical membrane of the proximal tubule, mediating urate reabsorption. In a Japanese population, the rs17300741 polymorphism in the SLC22A11 gene was associated with renal underexcretion type gout48).

NPT1 (SLC17A1) and NPT4 (SLC17A3) may be key players in renal urate secretion because SNPs in NPT1 and NPT4 weakly to moderately correlated with altered uric acid levels17). Gain-of-function variants of SLC17A1 may protect against gout, and loss-of-function mutations of SLC17A3 may be associated with hyperuricemia or gout10).

Multidrug-resistant proteins MRP2 (ABCC2) and MRP4 (ABCC4) transport urate in vitro, but there is no evidence at present for such a role in the kidney in vivo1). SNPs in these genes generally have a weak association with hyperuricemia17).

URAT1 and GLUT9 as Drug Targets

Pharmacologic agents for urate-lowering therapy can be classified into XO inhibitors, uricosuric agents, and uricases. Among these, uricosuric agents may act on urate transporters in the proximal tubule to increase urinary urate excretion. Both the inhibition of urate-reabsorbing transporters and stimulation of urate-secreting transporters might be conceivable, but the former is the known mechanism of action of uricosuric agents.

The apical URAT1 is the main target of uricosuric action. Probenecid and benzbromarone remarkably inhibit urate transport by URAT. Basolateral GLUT9 is also downregulated by probenecid and benzbromarone49). Lesinurad is a recently introduced uricosuric drug that may be used in combination with allopurinol or febuxostat. Its mechanism of action derives from the inhibition of URAT1, although OAT4 inhibition could also be associated3).

Losartan has a uricosuric action because of the inhibition of URAT1 and GLUT98). Sodium-glucose cotransporter-2 (SGLT2) inhibitors also have a uricosuric effect, probably induced by the inhibition of GLUT9b in the collecting duct50). On the other hand, pyrazinamide may induce hyperuricemia via the stimulation of URAT11).

Regulation of Urate Transporters in the Gut and Kidney

Serum uric acid levels are different between men and women, and the prevalence of hyperuricemia in adult males is higher than in females51). Sex hormones may play a role in regulating the expression or activity of urate transporters because SLC2A9 may have a stronger association with lower serum uric acid levels in females, whereas ABCG2 may have a stronger association with higher serum uric acid levels in males52). The ABCG2 Q140K+/+ knock-in mouse model is orthologous to the human ABCG2 Q141K variant. The Q140K+/+ mice showed hyperuricemia in males but not in females. Urinary urate excretion decreased in association with reduced renal ABCG2 protein abundance in male Q140K+/+ mice. However, urinary urate excretion and renal ABCG2 protein abundance were unaltered in female Q140K mice53).

The regulation of renal urate transporter expression by female hormones was investigated using ovariectomized mice with or without hormone replacement. Takiue et al. reported that estradiol suppressed the protein levels of URAT1, GLUT9, and ABCG2 in mouse kidneys54). On the other hand, the effect of testosterone was investigated in orchiectomized mice with or without testosterone replacement. Testosterone enhanced mRNA and protein levels of sodium-coupled monocarboxylate transporter 1 while those of GLUT9 were attenuated. Although the mRNA level of URAT1 was enhanced by testosterone, the corresponding levels of URAT1 protein remained unaffected55).

Other hormones might affect urate transporters in the kidney and gut for the regulation of urate homeostasis. Glucocorticoid was reported to increase urate excretion in mice by downregulating URAT156). Hyperuricemia may be associated with secondary hyperparathyroidism in patients with CKD. Sugimoto et al. reported that ABCG2 was downregulated by parathyroid hormone, increasing serum urate levels and these effects could be prevented by administration of the calcimimetic cinacalcet57).

Kim et al. investigated responses of urate transporters to increased uric acid intake in rat kidneys. Whereas URAT1 protein abundance was not affected, OAT1 protein abundance was increased by uric acid supplementation58). The up-regulation of OAT1 would exert stimulation of urinary urate excretion and might contribute to protection from hyperuricemia.

The oral administration of oxonic acid can induce hyperuricemia in rats. In this rat model, Nagura et al. found an increase in GLUT9 mRNA expression but no changes in URAT1 and ABCG2 mRNA in the kidney. Interestingly, ABCG2 mRNA increased in the ileum of rats with oxonic acid-induced hyperuricemia59). The lack of valid antibodies for urate transporters has limited studies at the protein level.

Because ABCG2 is localized in both the kidney and gut, the relative contribution of intestinal urate secretion will be increased in patients with CKD. Consistent with this, Yano et al. showed that ABCG2 mRNA expression increased in the ileum of rats with 5/6 nephrectomy60). In CKD, the role of intestinal urate transporters would be more important for maintaining urate homeostasis.

CONCLUSION

The kidney plays an important role in the maintenance of urate homeostasis because renal excretion of uric acid represents 70-80% of total uric acid excretion from the body. Approximately 10% of the glomerular filtered urate is normally excreted in urine after reabsorption and secretion in the proximal tubule. URAT1 (SLC22A12) is the major apical pathway for urate reabsorption, and GLUT9 (SLC2A9) is the principal pathway of basolateral urate exit from the proximal tubule cell in the human kidney. The loss-of-function mutation in either URAT1 or GLUT9 leads to renal hypouricemia. ABCG2, an apical ATP-driven efflux pump, functions in urate secretion by the proximal tubule and intestine, and ileal ABCG2 may be upregulated in CKD when renal urate transporters are downregulated. Genetic variations in SLC2A9, ABCG2, and SLC22A12 are associated with serum uric acid levels or gout (Fig. 3). Regulation of urate homeostasis at the level of urate transporters needs to be elucidated.

Fig. 3
A genome-wide association analysis of serum uric acid in 6,881 Korean individuals.
Manhattan plot depicts that the major urate transporter genes SLC22A12, ABCG2, and SLC2A9 are highly associated with hypouricemia or hyperuricemia. Adapted from Reference 40.

Notes

Conflict of Interest:The author declares no relevant financial interests

Acknowledgments

I thank Drs. Sua Kim and Chor ho Jo for their contribution to our work. This paper is based on a talk given by G.-H.K. at the annual meeting of KSN-IAC 2021 (May 8) in Seoul, Korea.

References

    1. Bobulescu IA, Moe OW. Renal transport of uric acid: evolving concepts and uncertainties. Adv Chronic Kidney Dis 2012;19:358–371.
    1. Oda M, Satta Y, Takenaka O, Takahata N. Loss of urate oxidase activity in hominoids and its evolutionary implications. Mol Biol Evol 2002;19:640–653.
    1. Benn CL, Dua P, Gurrell R, Loudon P, Pike A, Storer RI, Vangjeli C. Physiology of hyperuricemia and urate-lowering treatments. Front Med(Lausanne) 2018;5:160
    1. Bell S, Kolobova I, Crapper L, Ernst C. Lesch-Nyhan syndrome: models, theories, and therapies. Mol Syndromol 2016;7:302–311.
    1. Mandal AK, Mount DB. The molecular physiology of uric acid homeostasis. Annu Rev Physiol 2015;77:323–345.
    1. Richette P, Bardin T. Gout. Lancet 2010;375:318–328.
    1. So A, Thorens B. Uric acid transport and disease. J Clin Invest 2010;120:1791–1799.
    1. Enomoto A, Kimura H, Chairoungdua A, Shigeta Y, Jutabha P, Cha SH, Hosoyamada M, Takeda M, Sekine T, Igarashi T, Matsuo H, Kikuchi Y, Oda T, Ichida K, Hosoya T, Shimokata K, Niwa T, Kanai Y, Endou H. Molecular identification ofa renal urate anion exchanger that regulates blood urate levels. Nature 2002;417:447–452.
    1. Bahn A, Hagos Y, Reuter S, Balen D, Brzica H, Krick W, Burckhardt BC, Sabolic I, Burckhardt G. Identification of a new urate and high affinity nicotinate transporter, hOAT10 (SLC22A13). J Biol Chem 2008;283:16332–16341.
    1. Estiverne C, Mandal AK, Mount DB. Molecular pathophysiology of uric acid homeostasis. Semin Nephrol 2020;40:535–549.
    1. Xu X, Li C, Zhou P, Jiang T. Uric acid transporters hiding in the intestine. Pharm Biol 2016;54:3151–3155.
    1. Hagos Y, Stein D, Ugele B, Burckhardt G, Bahn A. Human renal organic anion transporter 4 operates as an asymmetric urate transporter. J Am Soc Nephrol 2007;18:430–439.
    1. Price KL, Sautin YY, Long DA, Zhang L, Miyazaki H, Mu W, Endou H, Johnson RJ. Human vascular smooth muscle cells express a urate transporter. J Am Soc Nephrol 2006;17:1791–1795.
    1. Augustin R, Carayannopoulos MO, Dowd LO, Phay JE, Moley JF, Moley KH. Identification and characterization of human glucose transporter-like protein-9 (GLUT9): alternative splicing alters trafficking. J Biol Chem 2004;279:16229–16236.
    1. Kimura T, Takahashi M, Yan K, Sakurai H. Expression of SLC2A9 isoforms in the kidney and their localization in polarized epithelial cells. PLoS One 2014;9:e84996
    1. Preitner F, Bonny O, Laverrière A, Rotman S, Firsov D, Da Costa A, Metref S, Thorens B. Glut9 is a major regulator of urate homeostasis and its genetic inactivation induces hyperuricosuria and urate nephropathy. Proc Natl Acad Sci USA 2009;106:15501–15506.
    1. Nigam SK, Bhatnagar V. The systems biology of uric acid transporters: the role of remote sensing and signaling. Curr Opin Nephrol Hypertens 2018;27:305–313.
    1. Huls M, Brown CD, Windass AS, Sayer R, van den Heuvel JJ, Heemskerk S, Russel FG, Masereeuw R. The breast cancer resistance protein transporter ABCG2 is expressed in the human kidney proximal tubule apical membrane. Kidney Int 2008;73:220–225.
    1. Hosomi A, Nakanishi T, Fujita T, Tamai I. Extra-renal elimination of uric acid via intestinal efflux transporter BCRP/ABCG2. PLoS One 2012;7:e30456
    1. Iharada M, Miyaji T, Fujimoto T, Hiasa M, Anzai N, Omote H, Moriyama Y. Type 1 sodium-dependent phosphate transporter (SLC17A1 Protein) is a Cl--dependent urate exporter. J Biol Chem 2010;285:26107–26113.
    1. Jutabha P, Anzai N, Kitamura K, Taniguchi A, Kaneko S, Yan K, Yamada H, Shimada H, Kimura T, Katada T, Fukutomi T, Tomita K, Urano W, Yamanaka H, Seki G, Fujita T, Moriyama Y, Yamada A, Uchida S, Wempe MF, Endou H, Sakurai H. Human sodium phosphate transporter 4 (hNPT4/ SLC17A3) as a common renal secretory pathway for drugs and urate. J Biol Chem 2010;285:35123–35132.
    1. Wright AF, Rudan I, Hastie ND, Campbell H. A ‘complexity’ of urate transporters. Kidney Int 2010;78:446–452.
    1. Togawa N, Miyaji T, Izawa S, Omote H, Moriyama Y. A Na+-phosphate cotransporter homologue (SLC17A4 protein) is an intestinal organic anion exporter. Am J Physiol Cell Physiol 2012;302:C1652–C1660.
    1. Enomoto A, Takeda M, Shimoda M, Narikawa S, Kobayashi Y, Kobayashi Y, Yamamoto T, Sekine T, Cha SH, Niwa T, Endou H. Interaction of human organic anion transporters 2 and 4 with organic anion transport inhibitors. J Pharmacol Exp Ther 2002;301:797–802.
    1. Henjakovic M, Hagos Y, Krick W, Burckhardt G, Burckhardt BC. Human organic anion transporter 2 is distinct from organic anion transporters 1 and 3 with respect to transport function. Am J Physiol Renal Physiol 2015;309:F843–F851.
    1. Mancikova A, Krylov V, Hurba O, Sebesta I, Nakamura M, Ichida K, Stiburkova B. Functional analysis of novel allelic variants in URAT1 and GLUT9 causing renal hypouricemia type 1 and 2. Clin Exp Nephrol 2016;20:578–584.
    1. Ichida K, Hosoyamada M, Hisatome I, Enomoto A, Hikita M, Endou H, Hosoya T. Clinical and molecular analysis of patients with renal hypouricemia in Japan-influence of URAT1 gene on urinary urate excretion. J Am Soc Nephrol 2004;15:164–173.
    1. Cheong HI, Kang JH, Lee JH, Ha IS, Kim S, Komoda F, Sekine T, Igarashi T, Choi Y. Mutational analysis of idiopathic renal hypouricemia in Korea. Pediatr Nephrol 2005;20:886–890.
    1. Shen H, Feng C, Jin X, Mao J, Fu H, Gu W, Liu A, Shu Q, Du L. Recurrent exercise-induced acute kidney injury by idiopathic renal hypouricemia with a novel mutation in the SLC2A9 gene and literature review. BMC Pediatr 2014;14:73
    1. Proctor P. Similar functions of uric acid and ascorbate in man? Nature 1970;228:868.
    1. Spitsin SV, Scott GS, Mikheeva T, Zborek A, Kean RB, Brimer CM, Koprowski H, Hooper DC. Comparison of uric acid and ascorbic acid in protection against EAE. Free Radic Biol Med 2002;33:1363–1371.
    1. Sautin YY, Johnson RJ. Uric acid: the oxidant-antioxidant paradox. Nucleosides Nucleotides Nucleic Acids 2008;27:608–619.
    1. Shimizu Y, Wakabayashi K, Totsuka A, Hayashi Y, Nitta S, Hara K, Akira M, Tomino Y, Suzuki Y. Exercise-induced acute kidney injury in a police officer with hereditary renal hypouricemia. Case Rep Nephrol Dial 2019;9:92–101.
    1. Zou Y, Du J, Zhu Y, Xie X, Chen J, Ling G. Associations between the SLC22A12 gene and gout susceptibility: a meta-analysis. Clin Exp Rheumatol 2018;36:442–447.
    1. Li C, Yu Q, Han L, Wang C, Chu N, Liu S. The hURAT1 rs559946 polymorphism and the incidence of gout in Han Chinese men. Scand J Rheumatol 2014;43:35–42.
    1. Duong NT, Ngoc NT, Thang NTM, Phuong BTH, Nga NT, Tinh ND, Quynh DH, Ton ND, Hai NV. Polymorphisms of ABCG2 and SLC22A12 genes associated with gout risk in Vietnamese population. Medicina (Kaunas) 2019;55:8
    1. Pavelcova K, Bohata J, Pavlikova M, Bubenikova E, Pavelka K, Stiburkova B. Evaluation of the influence of genetic variants of SLC2A9 (GLUT9) and SLC22A12 (URAT1) on the development of hyperuricemia and gout. J Clin Med 2020;9:2510
    1. Cho SK, Kim S, Chung JY, Jee SH. Discovery of URAT1 SNPs and association between serum uric acid levels and URAT1. BMJ Open 2015;5:e009360
    1. Hollis-Moffatt JE, Gow PJ, Harrison AA, Highton J, Jones PB, Stamp LK, Dalbeth N, Merriman TR. The SLC2A9 nonsynonymous Arg265His variant and gout: evidence for a population-specific effect on severity. Arthritis Res Ther 2011;13:R85
    1. Cho SK, Kim B, Myung W, Chang Y, Ryu S, Kim HN, Kim HL, Kuo PH, Winkler CA, Won HH. Polygenic analysis of the effect of common and low-frequency genetic variants on serum uric acid levels in Korean individuals. Sci Rep 2020;10:9179
    1. Vitart V, Rudan I, Hayward C, Gray NK, Floyd J, Palmer CN, Knott SA, Kolcic I, Polasek O, Graessler J, Wilson JF, Marinaki A, Riches PL, Shu X, Janicijevic B, Smolej-Narancic N, Gorgoni B, Morgan J, Campbell S, Biloglav Z, Barac-Lauc L, Pericic M, Klaric IM, Zgaga L, Skaric-Juric T, Wild SH, Richardson WA, Hohenstein P, Kimber CH, Tenesa A, Donnelly LA, Fairbanks LD, Aringer M, McKeigue PM, Ralston SH, Morris AD, Rudan P, Hastie ND, Campbell H, Wright AF. SLC2A9 is a newly identified urate transporter influencing serum urate concentration, urate excretion and gout. Nat Genet 2008;40:437–442.
    1. Zhang X, Yang X, Wang M, Li X, Xia Q, Xu S, Xu J, Cai G, Wang L, Xin L, Zou Y, Pan F. Association between SLC2A9 (GLUT9) gene polymorphisms and gout susceptibility: an updated meta-analysis. Rheumatol Int 2016;36:1157–1165.
    1. Lee YH, Seo YH, Kim JH, Choi SJ, Ji JD, Song GG. Associations between SLC2A9 polymorphisms and gout susceptibility: A meta-analysis. Z Rheumatol 2017;76:64–70.
    1. Testa A, Mallamaci F, Spoto B, Pisano A, Sanguedolce MC, Tripepi G, Leonardis D, Zoccali C. Association of a polymorphism in a gene encoding a urate transporter with CKD progression. Clin J Am Soc Nephrol 2014;9:1059–1065.
    1. Ichida K, Matsuo H, Takada T, Nakayama A, Murakami K, Shimizu T, Yamanashi Y, Kasuga H, Nakashima H, Nakamura T, Takada Y, Kawamura Y, Inoue H, Okada C, Utsumi Y, Ikebuchi Y, Ito K, Nakamura M, Shinohara Y, Hosoyamada M, Sakurai Y, Shinomiya N, Hosoya T, Suzuki H. Decreased extra-renal urate excretion is a common cause of hyperuricemia. Nat Commun 2012;3:764
    1. Matsuo H, Takada T, Ichida K, Nakamura T, Nakayama A, Ikebuchi Y, Ito K, Kusanagi Y, Chiba T, Tadokoro S, Takada Y, Oikawa Y, Inoue H, Suzuki K, Okada R, Nishiyama J, Domoto H, Watanabe S, Fujita M, Morimoto Y, Naito M, Nishio K, Hishida A, Wakai K, Asai Y, Niwa K, Kamakura K, Nonoyama S, Sakurai Y, Hosoya T, Kanai Y, Suzuki H, Hamajima N, Shinomiya N. Common defects of ABCG2, a high-capacity urate exporter, cause gout: a function-based genetic analysis in a Japanese population. Sci Transl Med 2009;1:5ra11.
    1. Son CN, Bang SY, Kim SH, Sung YK, Bae SC, Jun JB. ABCG2 Polymorphism Is Associated with Hyperuricemia in a Study of a Community-Based Korean Cohort. J Korean Med Sci 2017;32:1451–1459.
    1. Sakiyama M, Matsuo H, Shimizu S, Nakashima H, Nakayama A, Chiba T, Naito M, Takada T, Suzuki H, Hamajima N, Ichida K, Shimizu T, Shinomiya N. A common variant of organic anion transporter 4 (OAT4/SLC22A11) gene is associated with renal underexcretion type gout. Drug Metab Pharmacokinet 2014;29:208–210.
    1. Sattui SE, Gaffo AL. Treatment of hyperuricemia in gout: current therapeutic options, latest developments and clinical implications. Ther Adv Musculoskelet Dis 2016;8:145–159.
    1. Chino Y, Samukawa Y, Sakai S, Nakai Y, Yamaguchi J, Nakanishi T, Tamai I. SGLT2 inhibitor lowers serum uric acid through alteration of uric acid transport activity in renal tubule by increased glycosuria. Biopharm Drug Dispos 2014;35:391–404.
    1. Koo BS, Jeong HJ, Son CN, Kim SH, Kim HJ, Kim GH, Jun JB. Distribution of serum uric acid levels and prevalence of hyper- and hypouricemia in a Korean general population of 172,970. Korean J Intern Med 2021;36:S264–S272.
    1. Halperin Kuhns VL, Woodward OM. Sex Differences in Urate Handling. Int J Mol Sci 2020;21:4269
    1. Hoque KM, Dixon EE, Lewis RM, Allan J, Gamble GD, PhippsGreen AJ, Halperin Kuhns VL, Horne AM, Stamp LK, Merriman TR, Dalbeth N, Woodward OM. The ABCG2 Q141K hyperuricemia and gout associated variant illuminates the physiology of human urate excretion. Nat Commun 2020;11:2767
    1. Takiue Y, Hosoyamada M, Kimura M, Saito H. The effect of female hormones upon urate transport systems in the mouse kidney. Nucleosides Nucleotides Nucleic Acids 2011;30:113–119.
    1. Hosoyamada M, Takiue Y, Shibasaki T, Saito H. The effect of testosterone upon the urate reabsorptive transport system in mouse kidney. Nucleosides Nucleotides Nucleic Acids 2010;29:574–579.
    1. Li G, Han L, Ma R, Saeed K, Xiong H, Klaassen CD, Lu Y, Zhang Y. Glucocorticoids Increase Renal Excretion of Urate in Mice by Downregulating Urate Transporter 1. Drug Metab Dispos 2019;47:1343–1351.
    1. Sugimoto R, Watanabe H, Ikegami K, Enoki Y, Imafuku T, Sakaguchi Y, Murata M, Nishida K, Miyamura S, Ishima Y, Tanaka M, Matsushita K, Komaba H, Fukagawa M, Otagiri M, Maruyama T. Down-regulation of ABCG2, a urate exporter, by parathyroid hormone enhances urate accumulation in secondary hyperparathyroidism. Kidney Int 2017;91:658–670.
    1. Kim S, Lee CH, Kang CM, Kim GH. Effects of increased uric acid intake on the abundance of urate-anion exchanger and organic anion transporter proteins in the rat kidney. Electrolyte Blood Press 2007;5:62–67.
    1. Nagura M, Tamura Y, Kumagai T, Hosoyamada M, Uchida S. Uric acid metabolism of kidney and intestine in a rat model of chronic kidney disease. Nucleosides Nucleotides Nucleic Acids 2016;35:550–558.
    1. Yano H, Tamura Y, Kobayashi K, Tanemoto M, Uchida S. Uric acid transporter ABCG2 is increased in the intestine of the 5/6 nephrectomy rat model of chronic kidney disease. Clin Exp Nephrol 2014;18:50–55.

Metrics
Share
Figures

1 / 3

PERMALINK