Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2017 Jul 10;7:4939. doi: 10.1038/s41598-017-04949-2

Key Role for the Organic Anion Transporters, OAT1 and OAT3, in the in vivo Handling of Uremic Toxins and Solutes

Wei Wu 1,#, Kevin T Bush 2,#, Sanjay K Nigam 1,2,3,
PMCID: PMC5504054  PMID: 28694431

Abstract

In vitro data indicates that the kidney proximal tubule (PT) transporters of uremic toxins and solutes (e.g., indoxyl sulfate, p-cresol sulfate, kynurenine, creatinine, urate) include two “drug” transporters of the organic anion transporter (OAT) family: OAT1 (SLC22A6, originally NKT) and OAT3 (SLC22A8). Here, we have examined new and prior metabolomics data from the Oat1KO and Oat3KO, as well as newly obtained metabolomics data from a “chemical double” knockout (Oat3KO plus probenecid). This gives a picture of the in vivo roles of OAT1 and OAT3 in the regulation of the uremic solutes and supports the centrality of these “drug” transporters in independently and synergistically regulating uremic metabolism. We demonstrate a key in vivo role for OAT1 and/or OAT3 in the handling of over 35 uremic toxins and solutes, including those derived from the gut microbiome (e.g., CMPF, phenylsulfate, indole-3-acetic acid). Although it is not clear whether trimethylamine-N-oxide (TMAO) is directly transported, the Oat3KO had elevated plasma levels of TMAO, which is associated with cardiovascular morbidity in chronic kidney disease (CKD). As described in the Remote Sensing and Signaling (RSS) Hypothesis, many of these molecules are involved in interorgan and interorganismal communication, suggesting that uremia is, at least in part, a disorder of RSS.

Introduction

Uremic toxins/retention solutes comprise a large set of molecules that are normally cleared and excreted by the kidney, but which accumulate in the plasma of patients with renal insufficiency1. The accumulation of these endogenous compounds in the plasma (or failure in their excretion) can lead to the manifestation of uremic syndrome affecting the functions of other organs which, if untreated in its final stages, can be fatal1.

Uremic toxins have been divided into the following subgroups of compounds: (1) small water-soluble, non-protein-bound (e.g., urea); (2) small protein-bound (e.g., indoxyl sulfate); and (3) middle molecules (mainly small and large peptides)1, 2. Among these, the importance of water-soluble and protein-bound toxins has received considerable attention over the past several years1. Many of these are small, charged molecules whose uptake from the plasma and excretion via the kidney is mediated by transport proteins expressed on proximal tubule (PT) cells35. Among these are the multispecific transporters, organic anion transporter 1 (OAT1, SLC22A6, originally NKT6) and its close homolog OAT3 (SLC22A8)7, which are abundantly expressed on the basolateral membrane of PT cells (as well as several other epithelial tissues) and which together represent two of the most important transporters involved in the renal uptake and excretion of a wide variety of drugs, exogenous toxins, nutrients, and endogenous metabolites, including uremic toxins5, 8, 9.

OAT1 and OAT3 have previously been shown to interact with many of these uremic toxins and solutes in vitro and considerable in vivo data indicate that the OATs also handle many endogenous metabolites, including a number of uremic toxins and/or retention solutes5, 810. For example, previous targeted and untargeted limited metabolomics analyses of the Oat1KO and Oat3KO animals, albeit limited in that some were performed with older methods, instruments and databases, revealed significant alterations (up in plasma, down in urine) in the concentration of several uremic toxins and/or uremic retention solutes, including indoxyl sulfate, kynurenine, creatinine and urate1113. Since it is well-established that both the Oat1KO and Oat3KO animals display the expected defects in the handling of drugs and toxins12, 1422, such alterations in the plasma and/or urine concentrations of endogenous metabolites in the Oat1KO and Oat3KO animals are expected to reflect changes in their handling due to the absence of the transporters. Here, using a more comprehensive metabolomics analysis, levels of over 90 reported uremic toxins and uremic retention solutes were analyzed from the plasma of Oat1KO and Oat3KO animals, as well as from Oat3KO mice treated with probenecid, an inhibitor of OAT-mediated transport (“chemical double knockouts”)23.

Results

Despite clearly functioning as drug transporters, considerable in vitro and in vivo data indicate that OAT1 and OAT3 also handle many endogenous metabolites, including a number of uremic toxins and/or retention solutes5, 810, 24. In this study, levels of uremic toxins and uremic solutes were analyzed from the plasma of the Oat1KO and Oat3KO, as well as from Oat3KOs treated with probenecid (“chemical double knockouts”). Among the more than 600 metabolites examined, ~90 of them are water-soluble or protein-bound uremic toxins/retentions solutes identified by a number of groups (Supplemental Table S1)1, 2, 2529. Partial least squares discriminant analysis of these uremic toxins/retention solutes revealed separation of both Oat1KO and Oat3KO groups from their wildtype controls (Supplemental Figure 1).

While adding a few molecules which accumulate in the plasma of the Oat1KO (e.g., cysteine, kynurenate and S-adenosylhomocysteine), the metabolomics analysis of the plasma from the Oat1KO largely agreed with previous studies (Table 1)12, 13. For example, kynurenine, methionine and orotate, whose concentration was found to be altered in previous metabolomics analyses12, 13, also accumulated in the plasma of the Oat1KOs studied here (Table 1). On the other hand, analysis of the Oat3KO revealed considerably more alterations than previous attempts30, including significant increases in the plasma concentration of numerous uremic toxins/retention solutes, among them trimethylamine N-oxide (TMAO), indoxyl sulfate, p-cresol sulfate and 3-carboxy-4-methyl-5-propyl-2-furanpropanoate (CMPF) (Table 1). Taken together with in vitro data (Supplemental Table S2), the new in vivo metabolomics data showing the plasma accumulation of a number of uremic toxins provide strong support for the notion that Oat1 and Oat3 independently play an important role in the uptake and handling of a wide variety of water soluble and protein-bound uremic toxins/retention solutes12, 13, 30, 31. Moreover, certain compounds (e.g., kynurenine, indolelactate, indoxyl sulfate, methionine, creatinine, p-cresol sulfate and putrescine) appear to be in vivo substrates for both OATs.

Table 1.

Uremic toxins/retention solutes with significantly altered concentrations in Oat1 and Oat3 knockout mice based on previous and current metabolomics studies. aOAT1 references for previous studies: PMID:16354673; PMID:2146605). bOAT3 references for prior studies (PMID:18270321; PMID:18508962; PMID:2238083; PMID:2390220). cThe number in parentheses is the fold increase in plasma concentration for the listed metabolite. *p ≤ 0.05 for all fold increases except those for SAH and CMPF (p<0.1).

Previous Studies Additional Metabolites Found in This Study
OAT1 a OAT3 b OAT1 c OAT3 c
N2,N2-Dimethylguanosine Creatinine Cysteine (2.3) 1-Methyl-imidazoleacetate (1.5)
N-Methyladenosine Urate Kynurenate (1.6) 2-Aminophenol sulfate (3.2)
Indolelactate SAH (2.6)* 2-Oxindole-3-acetate (3.9)
Indoxyl sulfate CMPF (2.6)*
Kynurenine N-Acetyltryptophan (2.7)
Methionine Catechol sulfate (3.9)
Orotate Citrulline (1.3)
Phenylsulfate Imidazole propionate (4.5)
p-hydroxy-phenyllactic acid Indoleacetate (2.8)
Urate Indolelactate (4.6)
Xanthurenate Indoxyl sulfate (3.8)
Mannitol/sorbitol (1.7)
p-Cresol sulfate (3.8)
Phenylsulfate (2.8)
TMAO (4.5)

However, the most significant increases in plasma concentration of uremic toxins were observed in the Oat3KOs treated with probenecid—a so-called “chemical double knockout” (Tables 2 and 3). Thus far it has not been possible to create an OAT1/OAT3 double knockout mouse in part because of the chromosomal proximity of these two genes which essentially precludes recombination events in a genetic cross between the Oat1KO and Oat3KO. While probenecid does not exclusively inhibit OAT1 and OAT3 in vitro, by treating the Oat3KO with this drug, at least for organic anion transport across the basolateral membrane of the renal PT cell, a “chemical double knockout” was essentially created. Since, in the Oat3KO, the basolateral transport of organic anions by PT cells is dependent upon the remaining OAT (i.e., OAT1), treatment with probenecid should inhibit basolateral organic anion uptake mediated by OAT1 and lead to the accumulation of these molecules in the plasma, potentially allowing for identification of the uremic toxins with which the transporters interact.

Table 2.

Uremic Toxins/Retention Solutes Accumulating in the Plasma of Chemical Double Knockout (Oat3KO Treated with Probenecid versus Untreated Oat3KO).

Metabolite HMDB ID KEGG ID Fold Change (up in plasma) P-value
Xanthurenate HMDB00881 C02470 4.96 p ≤ 0.05
Kynurenate HMDB00715 C01717 4.85 p ≤ 0.05
Xanthosine HMDB00299 C01762 4.11 p ≤ 0.05
S-adenosylhomocysteine HMDB00939 C00021 3.57 p ≤ 0.05
Homovanillate sulfate HMDB11719 3.54 p ≤ 0.05
Isovalerylglycine HMDB00678 3.25 p ≤ 0.05
4-Hydroxyhippurate HMDB13678 2.83 p ≤ 0.05
Kynurenine HMDB00684 C00328 2.25 p ≤ 0.05
Putrescine HMDB01414 C00134 2.04 p ≤ 0.05
Spermidine HMDB01257 C00315 1.85 p ≤ 0.05
5-methylthioadenosine HMDB01173 C00170 1.80 p ≤ 0.05
Cysteine HMDB00574 C00097 1.71 p ≤ 0.05
Hypoxanthine HMDB00157 C00262 1.70 p ≤ 0.05
Orotate HMDB00226 C00295 1.70 p ≤ 0.05
p-Hydroxy-phenyllactate HMDB00755 C03672 1.50 p ≤ 0.05
Urate HMDB00289 C00366 1.40 p ≤ 0.05
Threonate HMDB00943 C01620 1.37 p ≤ 0.05
Dimethylglycine HMDB00092 C01026 1.30 p ≤ 0.05
Creatinine HMDB00562 C00791 1.26 p ≤ 0.05
p-Cresol sulfate HMDB11635 C01468 3.80 0.05 < p < 0.10

Fold Change in plasma concentration of uremic toxins/retention solutes [(Oat3KO + probenecid) vs (Oat3KO untreated control)].

Table 3.

Uremic Toxins and/or Retention Solutes Interactions with Oat1KO, Oat3KO, or Chemical Double Knockout.

Metabolite In Vivo Metabolomics *Existence of In Vitro Support for Interaction with Transporter
Oat1KO Compared to Wildtype Oat3KO Compared to Wildtype Chemical Double KO (Oat3KO + Probenecid) Compared to Oat3KO OAT1 OAT3
N6-Methyladenosine Down in urine (A)
Phenylsulfate Up in plasma (A)
N2,N2-dimethyl-guanosine Down in urine (A)
Methionine Up in plasma (A)
Xanthurenate Down in urine (A) Up in plasma (this study)
Orotate Down in urine (A, B) Up in plasma (this study)
p-Hydroxy-phenyllactate Up in plasma (A) Up in plasma (this study)
S-Adenosylhomocysteine Up in plasma; this study Up in plasma (this study)
Cysteine Up in plasma; this study Up in plasma (this study)
Kynurenate Up in plasma; this study Up in plasma (this study)
Uric acid Down in urine (A) Decreased renal secretion (C) Down in urine (B) Up in plasma (this study)
Creatinine Decreased renal secretion (D) Decreased renal secretion (D) Up in plasma (this study)
Kynurenine Up in plasma (A) Up in plasma (this study) Up in plasma (this study)
Indolelactate Up in plasma (A) Up in plasma (this study)
Indoxyl sulfate Up in plasma (A) Up in plasma (this study)
2-Aminophenol sulfate Up in plasma (this study)
2-Oxindole-3-acetate Up in plasma (this study)
Catechol sulfate Up in plasma (this study)
N-Acetyltryptophan Up in plasma (this study)
Mannitol/Sorbitol Up in plasma (this study)
Trimethylamine-N-oxide Up in plasma (this study)
1-Methylimidazoleacetate Up in plasma (this study)
Citrulline Up in plasma (this study)
3-carboxy-4-methyl-5-propyl-2-furanpropionate Up in plasma (this study)
Imidazole propionate Up in plasma (this study)
Indoleacetate Up in plasma (this study)
p-Cresol sulfate Up in plasma (this study) Up in plasma (this study)
Putrescine Up in plasma (this study) Up in plasma (this study)
4-Hydroxyhippurate Up in plasma (this study)
5-Methylthioadenosine Up in plasma (this study)
Dimethylglycine Up in plasma (this study)
Homovanillate sulfate Up in plasma (this study)
Hypoxanthine Up in plasma (this study)
Isovaleryglycine Up in plasma (this study)
Spermidine Up in plasma (this study)
Threonate Up in plasma (this study)
Xanthosine Up in plasma (this study)

(A−D) Columns 2–3 indicate findings from previous studies: A–PMID: 2147660, B–PMID:16354673; C – PMID: 18270321; D–PMID: 22338083. (−) Columns 2–4 indicate no significant change detected in plasma concentration of the indicated metabolite in the in vivo metabolomics analyses. Blanks in Columns 5–6 indicate that in vitro uptake/inhibition studies have not been performed for the indicated metabolite.

*See Supplemental Table 2 for Km, Ki, IC50 and PMIDs.

As above, partial least squares discriminant analysis of uremic toxins/retention solutes revealed separation between the Oat3KO and probenecid-treated-Oat3KO groups based on the metabolite profile of the uremic toxins/retention solutes (Fig. 1) and analysis of the plasma samples derived from these “chemical double knockout” models revealed profound plasma accumulation of several known and/or suspected uremic toxins/solutes compared to the untreated Oat3KO (Table 2). For example, kynurenate was only slightly elevated in the plasma of Oat3KO mice, however, following treatment of the Oat3KO with probenecid, the plasma level of this uremic toxin significantly increased almost 5-fold compared to that seen in the Oat3KO (Table 2). This accounted for nearly 80% of the total fold-increase in the plasma seen in the Oat3KO (i.e., total of fold-increase seen in both the Oat3KO and the probenecid-treated Oat3KO). Thus, since, as described above, probenecid is inhibiting OAT1-mediated basolateral uptake from the plasma in the proximal tubule of the Oat3KO, this data strongly suggests that the in vivo PT uptake and handling of kynurenate is largely mediated by OAT1 (Fig. 2). A similar phenomenon was seen with other toxins/retention solutes, including cysteine, S-adenosylhomocysteine, kynurenine, putrescine, spermidine, orotate and hypoxanthine (Tables 13). On the other hand, p-cresol sulfate appears to be handled more or less equally by both OAT1 and OAT3 (Fig. 2), with ~50% of the total fold increase in the plasma of the Oat3KO following treatment with probenecid, a finding that is in agreement with in vitro data (Supplemental Table 2). In contrast, while 3-carboxy-4-methyl-5-propyl-2-furanpropanoate (CMPF) and trimethylamine-N-oxide (TMAO) were both elevated in the plasma of Oat3KO mice, treatment of the Oat3KO with probenecid led to relatively small additional accumulations of these metabolites compared to the Oat3KO which were not statistically significant, indicating that the majority of the observed plasma accumulation of these metabolites in the Oat3KO is likely due to the absence of OAT3 (Fig. 2). This suggests that CMPF and TMAO are more likely to interact with OAT3 than OAT1.

Figure 1.

Figure 1

Partial least squares discriminant analysis of uremic toxins/retention solutes reveals separation between plasma metabolites from Oat3KO mice (KO3; red) and Oat3KO mice treated with probenecid (PROBKO3; green). The dots represent each individual uremic toxin/retention solute metabolite profile and the shaded areas represent the 95% confidence interval.

Figure 2.

Figure 2

Stacked bar graph of the overall fold change in the plasma levels of the Oat3KO for some uremic toxins [i.e., Xanthurenate, Kynurenate, Creatinine, p-Cresol sulfate, 3-carboxy-4-methyl-5-propyl-2-furanpropionate (CMPF), Indoxyl sulfate and Trimethylamine N-oxide (TMAO)]. The data represents the percentage of the total fold change in the overall plasma concentration for each of these toxins/retention solutes in the Oat3KO (grey boxes) and in the Oat3KO treated with probenecid (black boxes). As described in the text, since in the Oat3KO, treatment with probenecid inhibits OAT1-mediated renal uptake, this allows one to see the contribution of each organic anion transporter to the handling of these metabolites (please also see Table 3).

While data supports the notion that CMPF is more a substrate of OAT3 rather than OAT125, 32, recently published in vitro data seems to suggest that OAT3 does not directly transport TMAO (albeit this assay was performed at a single high concentration)25. This would seem to be consistent with TMAO’s more cationic character; however, an early physiological study found that the renal excretion of TMAO could be inhibited by probenecid, a finding which was, at that time, attributed to probenecid-mediated blockage of the movement of TMAO into the urine via organic anion transporters localized to apical membrane of PT cells33. Nevertheless, the finding here of increased plasma levels of TMAO in the Oat3KOs would appear to indicate that renal uptake of endogenous TMAO from the plasma is, directly or indirectly, mediated via this basolateral organic anion transporter. It is important to note that OAT3 has been shown to have a greater capacity than OAT1 to interact with cationic compounds, possibly including TMAO34, 35.

We have analyzed the in vivo roles of renal proximal tubule organic anion transporters, OAT1 and OAT3, individually and in combination, in mediating the uptake and excretion of uremic toxins and uremic retention solutes. Considerable new metabolomics data from the Oat1KO, Oat3KO and a “chemical double knockout” (Oat3KO treated with probenecid) when analyzed in the context of limited previous data clearly links the OATs to the handling of numerous uremic toxins/retention solutes. When considered in light of existing in vitro uremic toxin interaction data for these transporters (Supplemental Table 2), it is clear that OAT1 and OAT3 are required for effective renal elimination of many of the “best known” uremic toxins and retention solutes (Table 3).

Discussion

OAT1/SLC22A6 and OAT3/SLC22A8 are not only both expressed in the basolateral membrane of renal PT cells, but they share a high degree of sequence similarity. In fact, they are each other’s closest homolog and share a great deal of similar functionality35, 36. Nevertheless, the data from OAT knockout mice suggests that, while these transporters handle many uremic toxins, certain toxins/solutes prefer OAT1 (e.g., kynurenate, kynurenine, orotate), while others prefer OAT3 (e.g., TMAO, CMPF) with some appearing to be capable of interacting more or less equally with both (e.g., p-cresol sulfate) (Fig. 2; Tables 13). Therefore it was of interest to determine if there is an additive or synergistic effect of simultaneously blocking transport mediated by OAT1 and OAT3. Thus, in this study, a “chemical double” knockout was generated by treating the Oat3KO with probenecid.

For the majority of the metabolites altered in the Oat3KO, it seems that the additional inhibition of OAT1 by probenecid did not significantly alter the plasma levels of the toxin/solute, suggesting that, while the transporters might share some ligands, for the large part the transporters appear to have a significant level of substrate selectivity. Nevertheless, there were a group of toxins/solutes which only show a significant level of accumulation in the plasma when both transporters were inhibited (Table 3), suggesting that both of the transporters contribute to the uptake and elimination of these molecules.

Although we cannot absolutely exclude minor effects of probenecid on other transporters, such as URAT1 (RST) and MRPs, based on available data it seems likely that OAT1 is likely to be the major transporter affected by treatment of the Oat3KO with probenecid. For example, probenecid exhibits a significant inhibition potency towards OAT1 (and OAT3) with K is in the low micromolar range (i.e., 1.3–32 µM depending on the substrates and model system used), while the inhibition potency towards other transporters has been found to be much higher37. For example, the K i for OAT2 is 766 µM and MRP4 is >1000 µM, while that for the apical uptake transporter, OAT4, is ~55 µM and for the apical extrusion transporter, MRP2, it is ~45 µM37.

Furthermore, early pharmacokinetic studies in rodents demonstrated that the effective unbound plasma concentration of probenecid following a single intraperitoneal dose of probenecid at a concentration of 250 mg/kg was roughly equivalent to that of a human receiving a typical oral dose of 0.5–2.0 g of probenecid38, which is in the range of 3–50 µM39. This concentration is well within the range of that needed to inhibit OAT1 in the Oat3KO, but well below that needed to inhibit OAT2 or MRP4. Although it approaches micromolar concentrations for inhibiting OAT4 and MRP2, these transporters are found on the apical membrane of the proximal tubule cell and are involved in reabsorption from the tubular lumen. Since the endogenous metabolites we are discussing are those which are accumulating in the plasma of the Oat3KO treated with probenecid, the inhibition of these transporters is likely to be less of a concern.

Moreover, the likelihood that the changes we are observing are due to probenecid-mediated inhibition of organic anion transporters other than OAT1, such as URAT1 (RST) seems low since metabolomics studies of the knockout and extensive in vitro studies of URAT1 (RST) in relation to uremic toxins do not support the view that it transports any uremic toxin other than uric acid4, 40, and the many in vivo studies of OAT1 and OAT3 knockouts are quite consistent with in vivo probenecid data. Moreover, if basolateral OAT1 and OAT3 are blocked by a combination of gene deficiency and probenecid, apical extrusion transporters such as MRPs would not significantly come into play for organic anion uremic toxins in the blood.

Since both of these transporters mediate the pharmacokinetics of a wide variety of drugs, the data raises the possibility that drugs that compete for transport via OAT1 and OAT3 with the uremic toxins have the potential to induce alterations in the levels of the toxins41. For example, the accumulation of neurotransmitter metabolites and drugs in the brain in the setting of chronic kidney failure has been suggested to be due to reduced OAT3-mediated brain-to-blood transport due to the increased concentration of uremic toxins such as indoxyl sulfate42. Taken together, this could provide a partial explanation for some of the off-target effects of some drugs, including those associated with metabolic syndrome, particularly in the setting of chronic disease and long-term drug treatment5, 8, 9. For example, the furan fatty acid metabolite 3-carboxy-4-methyl-5-propyl-2-furanpropanoic acid (CMPF) is a uremic toxin whose renal elimination is mediated by the OATs. This metabolite has recently been shown to act on pancreatic β-cells leading to reduced insulin biosynthesis and represents a potential link to glucose intolerance32. Thus, one can see how competition for OAT-mediated clearance of CMPF and a drug could result in a drug-metabolite interaction at the level of the renal transporter affecting the concentrations, not only of the drug, but of the endogenous metabolite as well, which in the case of CMPF could lead to alterations in glucose tolerance with its effects on the pancreas.

In further support of this notion, it was recently demonstrated that administration of either ketoprofen or diclofenac (non-steroidal anti-inflammatory drugs (NSAIDs) capable of inhibiting OAT1- and OAT3-mediated transport), significantly decreased the renal clearance of indoxyl sulfate thereby increasing systemic exposure to this uremic toxin43. It was suggested that this NSAID-mediated increase in the concentration of indoxyl sulfate, might contribute to the progression of indoxyl sulfate-induced cardiovascular disease and, at least in part, explain the pathogenesis of analgesic nephropathy44. Taken together, the major point is that partial blockade of OAT1- or OAT3-mediated transport of uremic toxins by competing drugs has the potential to alter the concentrations of certain toxins and lead to the distal cascade effects of the toxin.

Our study also supports the relevance of the remote sensing and signaling hypothesis to the pathophysiology of uremia and CKD5, 8, 9, 4548. This hypothesis emphasizes the centrality of multispecific SLC and ABC transporters to inter-organ and inter-organismal small molecule communication throughout the body and thus the importance of the transporters to modulation of metabolism and signaling5, 8, 9, 30, 45, 47, 49, 50. In this regard, it is worth noting that some of the uremic toxins are derived from metabolites generated by the microbiome in the gut. For example, indoxyl sulfate, ultimately derived from metabolism of tryptophan by the gut bacteria, is absorbed from the intestine into the blood as indole, which is then sulfated in the liver to indoxyl sulfate which is transported into various organs and/or excreted by the kidney51. This amounts to interorganismal movement of a small organic anion capable of inducing toxicity or activating signaling pathways via nuclear receptor signaling. For example, indoxyl sulfate, kynurenine and kynurenate signal via the aryl hydrocarbon receptor (AHR)5254, which is expressed in most tissues, including those affected in uremia such as the liver, intestine and kidney55, 56. Kynurenate can also activate the orphan G-couple protein receptor, GPR3557, which is largely expressed in tissues affected in uremia, such as the gut and CNS58, 59. CMPF, on the other hand, has been implicated in a glucose sensing mechanism involving OAT3, which is expressed in the pancreas (among other non-renal tissues), leading to altered metabolism32. Of note, metabolic reconstructions of “omics” data from the Oat1KO implicated regulation by OAT1 and OAT3 of a number of biochemical pathways reported to be altered in uremia (e.g., polyamine and purine metabolism)11. Although much more work needs to be done in this regard, it is possible that high levels of uremic toxins inhibit the normal metabolic functions of the OATs via uremic toxin-metabolite competition at the level of the transporter; certainly there is evidence for this in the case of uremic toxin-drug competition25, 43.

The results support the roles of organic anion transporters in the handling of a number of uremic toxins/retention solutes that follow interorgan and interorganismal (gut microbiome-host) communication pathways that may include the activation of key signaling pathways affecting a wide variety of metabolic processes, including those which are known to be affected in uremic syndrome pathophysiology46. This general view is also supported by metabolic reconstructions of “omics” data from the Oat1KO and Oat3KO mice11, 30, 49, where it has been demonstrated that Oats regulate biochemical pathways of interorgan and interorganismal communication such as those involving gut microbiome metabolites, energy metabolism, purine metabolism, antioxidants and lipid metabolism and many other pathways altered in CKD. Taken together, this suggests that uremia is in part a disease of disordered remote sensing and signaling5, 8, 9, 11, 4547.

Methods

Animals

All experimental protocols were approved by The University of California San Diego Institutional Animal Care and Use Committee (IACUC). The animals were handled in accordance with the Institutional Guidelines on the Use of Live Animals for Research; all experiments involving the use of animals were also conducted in accordance with the Institutional Guidelines on the Use of Live Animals for Research. Adult (n = 3) wildtype, Oat1- and Oat3-deficient male mice received a single, daily intraperitoneal (i.p.) injection of either 200 mg/kg water-soluble probenecid (Invitrogen, Carlsbad, CA) [0.02 mg/µL in PBS (10 µL/g of body weight)] or PBS (sham-treated control) for three days. The animals were housed separately under a 12-h light-dark cycle and were provided access to food (standard diet) and water ad libitum. On day three, two hours before samples were taken, the mice were given a final i.p. injection of either probenecid or PBS. Two hours after this last treatment, blood was collected, and plasma was isolated and stored at −80 °C until being shipped for analysis as described below.

Metabolomic Analysis, Data Extraction, Compound Identification, Curation and Statistics

Individual, unpooled samples were measured by the Metabolon analytical system (Metabolon, Inc., Durham, NC)60, 61. Samples were prepared and subjected to ultrahigh performance liquid chromatography-tandem mass spectroscopy (UPLC-MS/MS) utilizing an ACQUITY ultra-performance liquid chromatography (UPLC) (Waters, Milford, MA) and a Q-Exactive high resolution/accurate mass spectrometer interfaced with a heated electrospray ionization (HESI-II) source and Orbitrap mass analyzer operated at 35,000 mass resolution (Thermo Scientific, Waltham, MA). Raw data was extracted, peak-identified and QC processed using Metabolon’s hardware and software60, 61. Two-way ANOVA testing was used to calculate the p-values and the metabolites that were selected for display in figures and tables had either: (1) a fold change ≥1.2 with a p-value ≤ 0.05; or (2) a fold change ≥2.0 with a p-value ≤ 0.1 in at least one of the various comparisons (e.g., Oat1KO vs WT; Oat3KO vs WT; Oat3KO + probenecid vs Oat3KO). The separability of the uremic toxins/retention solutes in the wildtype, Oat3KO, and probenecid-treated Oat3KO plasma samples was assessed by partial least squares discriminant analysis (PLS-DA) using Metaboanalyst 3.0 (http://www.metaboanalyst.ca/)62.

Electronic supplementary material

Supplemental Data (432.7KB, pdf)

Acknowledgements

This work was supported in whole or part by National Institutes of Health Grants R01-DK109392 and U54-HD090259.

Author Contributions

W.W. and K.T.B. contributed equally to the work. W.W. and K.T.B. performed data analysis and contributed to writing of the manuscript. S.K.N. conceived the hypothesis, supervised the experiments and analysis and edited the manuscript. All authors reviewed the manuscript.

Competing Interests

The authors declare that they have no competing interests.

Footnotes

W.W. and K.T.B. contributed equally to this work.

Electronic supplementary material

Supplementary information accompanies this paper at doi:10.1038/s41598-017-04949-2

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Vanholder RC, Eloot S, Glorieux GL. Future Avenues to Decrease Uremic Toxin Concentration. Am J Kidney Dis. 2016;67:664–676. doi: 10.1053/j.ajkd.2015.08.029. [DOI] [PubMed] [Google Scholar]
  • 2.Duranton F, et al. Normal and pathologic concentrations of uremic toxins. J Am Soc Nephrol. 2012;23:1258–1270. doi: 10.1681/ASN.2011121175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Lowenstein J, Grantham JJ. The rebirth of interest in renal tubular function. Am J Physiol Renal Physiol. 2016;310:F1351–1355. doi: 10.1152/ajprenal.00055.2016. [DOI] [PubMed] [Google Scholar]
  • 4.Suchy-Dicey AM, et al. Tubular Secretion in CKD. J Am Soc Nephrol. 2016;27:2148–2155. doi: 10.1681/ASN.2014121193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Nigam SK, et al. Handling of Drugs, Metabolites, and Uremic Toxins by Kidney Proximal Tubule Drug Transporters. Clin J Am Soc Nephrol. 2015;10:2039–2049. doi: 10.2215/CJN.02440314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Lopez-Nieto CE, et al. Molecular cloning and characterization of NKT, a gene product related to the organic cation transporter family that is almost exclusively expressed in the kidney. J Biol Chem. 1997;272:6471–6478. doi: 10.1074/jbc.272.10.6471. [DOI] [PubMed] [Google Scholar]
  • 7.Zhu, C. et al. Evolutionary Analysis and Classification of OATs, OCTs, OCTNs, and Other SLC22 Transporters: Structure-Function Implications and Analysis of Sequence Motifs. PLOS ONE10, e0140569, doi:10.1371/journal.pone.0140569 (2015). [DOI] [PMC free article] [PubMed]
  • 8.Nigam SK. What do drug transporters really do? Nat Rev Drug Discov. 2015;14:29–44. doi: 10.1038/nrd4461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Nigam SK, et al. The organic anion transporter (OAT) family: a systems biology perspective. Physiol Rev. 2015;95:83–123. doi: 10.1152/physrev.00025.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.VanWert AL, Gionfriddo MR, Sweet DH. Organic anion transporters: discovery, pharmacology, regulation and roles in pathophysiology. Biopharm Drug Dispos. 2010;31:1–71. doi: 10.1002/bdd.693. [DOI] [PubMed] [Google Scholar]
  • 11.Ahn SY, et al. Linkage of organic anion transporter-1 to metabolic pathways through integrated “omics”-driven network and functional analysis. J Biol Chem. 2011;286:31522–31531. doi: 10.1074/jbc.M111.272534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Eraly SA, et al. Decreased renal organic anion secretion and plasma accumulation of endogenous organic anions in OAT1 knock-out mice. J Biol Chem. 2006;281:5072–5083. doi: 10.1074/jbc.M508050200. [DOI] [PubMed] [Google Scholar]
  • 13.Wikoff WR, Nagle MA, Kouznetsova VL, Tsigelny IF, Nigam SK. Untargeted metabolomics identifies enterobiome metabolites and putative uremic toxins as substrates of organic anion transporter 1 (Oat1) J Proteome Res. 2011;10:2842–2851. doi: 10.1021/pr200093w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Eraly SA, et al. Multiple organic anion transporters contribute to net renal excretion of uric acid. Physiol Genomics. 2008;33:180–192. doi: 10.1152/physiolgenomics.00207.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Sweet DH, et al. Impaired organic anion transport in kidney and choroid plexus of organic anion transporter 3 (Oat3 [Slc22a8]) knockout mice. J Biol Chem. 2002;277:26934–26943. doi: 10.1074/jbc.M203803200. [DOI] [PubMed] [Google Scholar]
  • 16.Sykes D, et al. Organic anion transport in choroid plexus from wild-type and organic anion transporter 3 (Slc22a8)-null mice. Am J Physiol Renal Physiol. 2004;286:F972–978. doi: 10.1152/ajprenal.00356.2003. [DOI] [PubMed] [Google Scholar]
  • 17.Nagle MA, Wu W, Eraly SA, Nigam SK. Organic anion transport pathways in antiviral handling in choroid plexus in Oat1 (Slc22a6) and Oat3 (Slc22a8) deficient tissue. Neurosci Lett. 2013;534:133–138. doi: 10.1016/j.neulet.2012.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Torres AM, Dnyanmote AV, Bush KT, Wu W, Nigam SK. Deletion of multispecific organic anion transporter Oat1/Slc22a6 protects against mercury-induced kidney injury. J Biol Chem. 2011;286:26391–26395. doi: 10.1074/jbc.M111.249292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Truong DM, Kaler G, Khandelwal A, Swaan PW, Nigam SK. Multi-level analysis of organic anion transporters 1, 3, and 6 reveals major differences in structural determinants of antiviral discrimination. J Biol Chem. 2008;283:8654–8663. doi: 10.1074/jbc.M708615200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Vallon V, et al. Overlapping in vitro and in vivo specificities of the organic anion transporters OAT1 and OAT3 for loop and thiazide diuretics. Am J Physiol Renal Physiol. 2008;294:F867–873. doi: 10.1152/ajprenal.00528.2007. [DOI] [PubMed] [Google Scholar]
  • 21.Vanwert AL, Bailey RM, Sweet DH. Organic anion transporter 3 (Oat3/Slc22a8) knockout mice exhibit altered clearance and distribution of penicillin G. Am J Physiol Renal Physiol. 2007;293:F1332–1341. doi: 10.1152/ajprenal.00319.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Vanwert AL, Srimaroeng C, Sweet DH. Organic anion transporter 3 (oat3/slc22a8) interacts with carboxyfluoroquinolones, and deletion increases systemic exposure to ciprofloxacin. Mol Pharmacol. 2008;74:122–131. doi: 10.1124/mol.107.042853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Robbins N, Koch SE, Tranter M, Rubinstein J. The history and future of probenecid. Cardiovasc Toxicol. 2012;12:1–9. doi: 10.1007/s12012-011-9145-8. [DOI] [PubMed] [Google Scholar]
  • 24.Koepsell H. The SLC22 family with transporters of organic cations, anions and zwitterions. Mol Aspects Med. 2013;34:413–435. doi: 10.1016/j.mam.2012.10.010. [DOI] [PubMed] [Google Scholar]
  • 25.Hsueh CH, et al. Identification and Quantitative Assessment of Uremic Solutes as Inhibitors of Renal Organic Anion Transporters, OAT1 and OAT3. Mol Pharm. 2016;13:3130–3140. doi: 10.1021/acs.molpharmaceut.6b00332. [DOI] [PubMed] [Google Scholar]
  • 26.Liabeuf S, Drueke TB, Massy ZA. Protein-bound uremic toxins: new insight from clinical studies. Toxins (Basel) 2011;3:911–919. doi: 10.3390/toxins3070911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Niwa T. Targeting protein-bound uremic toxins in chronic kidney disease. Expert Opin Ther Targets. 2013;17:1287–1301. doi: 10.1517/14728222.2013.829456. [DOI] [PubMed] [Google Scholar]
  • 28.Rhee EP, et al. Metabolite profiling identifies markers of uremia. J Am Soc Nephrol. 2010;21:1041–1051. doi: 10.1681/ASN.2009111132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Tanaka H, Sirich TL, Plummer NS, Weaver DS, Meyer TW. An Enlarged Profile of Uremic Solutes. PLoS One. 2015;10:e0135657. doi: 10.1371/journal.pone.0135657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wu W, et al. Multispecific drug transporter Slc22a8 (Oat3) regulates multiple metabolic and signaling pathways. Drug Metab Dispos. 2013;41:1825–1834. doi: 10.1124/dmd.113.052647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Vallon V, et al. Organic anion transporter 3 contributes to the regulation of blood pressure. J Am Soc Nephrol. 2008;19:1732–1740. doi: 10.1681/ASN.2008020180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Prentice KJ, et al. The furan fatty acid metabolite CMPF is elevated in diabetes and induces beta cell dysfunction. Cell Metab. 2014;19:653–666. doi: 10.1016/j.cmet.2014.03.008. [DOI] [PubMed] [Google Scholar]
  • 33.Acara M, Camiolo S, Rennick B. Renal N-oxidation of trimethylamine in the chicken during tubular excretion. Drug Metab Dispos. 1977;5:82–90. [PubMed] [Google Scholar]
  • 34.Ahn SY, Eraly SA, Tsigelny I, Nigam SK. Interaction of organic cations with organic anion transporters. J Biol Chem. 2009;284:31422–31430. doi: 10.1074/jbc.M109.024489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Liu, H. C. et al. Analysis of Molecular Properties of Drugs Interacting with SLC22 Transporters OAT1, OAT3, OCT1, and OCT2: A Machine-Learning Approach. J Pharmacol Exp Ther, doi:10.1124/jpet.116.232660 (2016). [DOI] [PMC free article] [PubMed]
  • 36.Kaler G, et al. Structural variation governs substrate specificity for organic anion transporter (OAT) homologs. Potential remote sensing by OAT family members. J Biol Chem. 2007;282:23841–23853. doi: 10.1074/jbc.M703467200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Yin J, Wang J. Renal drug transporters and their significance in drug-drug interactions. Acta Pharm Sin B. 2016;6:363–373. doi: 10.1016/j.apsb.2016.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Dayton PG, Perel JM. The metabolism of probenecid in man. Ann N Y Acad Sci. 1971;179:399–402. doi: 10.1111/j.1749-6632.1971.tb46916.x. [DOI] [PubMed] [Google Scholar]
  • 39.Emanuelsson BM, Beermann B, Paalzow LK. Non-linear elimination and protein binding of probenecid. Eur J Clin Pharmacol. 1987;32:395–401. doi: 10.1007/BF00543976. [DOI] [PubMed] [Google Scholar]
  • 40.Eraly SA, Liu HC, Jamshidi N, Nigam SK. Transcriptome-based reconstructions from the murine knockout suggest involvement of the urate transporter, URAT1 (slc22a12), in novel metabolic pathways. Biochem Biophys Rep. 2015;3:51–61. doi: 10.1016/j.bbrep.2015.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Emami Riedmaier A, Nies AT, Schaeffeler E, Schwab M. Organic anion transporters and their implications in pharmacotherapy. Pharmacol Rev. 2012;64:421–449. doi: 10.1124/pr.111.004614. [DOI] [PubMed] [Google Scholar]
  • 42.Ohtsuki S, et al. Role of blood-brain barrier organic anion transporter 3 (OAT3) in the efflux of indoxyl sulfate, a uremic toxin: its involvement in neurotransmitter metabolite clearance from the brain. J Neurochem. 2002;83:57–66. doi: 10.1046/j.1471-4159.2002.01108.x. [DOI] [PubMed] [Google Scholar]
  • 43.Yu CP, et al. Effects of nonsteroidal anti-inflammatory drugs on the renal excretion of indoxyl sulfate, a nephro-cardiovascular toxin, in rats. Eur J Pharm Sci. 2017;101:66–70. doi: 10.1016/j.ejps.2017.02.007. [DOI] [PubMed] [Google Scholar]
  • 44.Yaxley J. Common Analgesic Agents and Their Roles in Analgesic Nephropathy: A Commentary on the Evidence. Korean J Fam Med. 2016;37:310–316. doi: 10.4082/kjfm.2016.37.6.310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ahn SY, Nigam SK. Toward a systems level understanding of organic anion and other multispecific drug transporters: a remote sensing and signaling hypothesis. Mol Pharmacol. 2009;76:481–490. doi: 10.1124/mol.109.056564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Saito H. Pathophysiological regulation of renal SLC22A organic ion transporters in acute kidney injury: pharmacological and toxicological implications. Pharmacol Ther. 2010;125:79–91. doi: 10.1016/j.pharmthera.2009.09.008. [DOI] [PubMed] [Google Scholar]
  • 47.Wu W, Dnyanmote AV, Nigam SK. Remote communication through solute carriers and ATP binding cassette drug transporter pathways: an update on the remote sensing and signaling hypothesis. Mol Pharmacol. 2011;79:795–805. doi: 10.1124/mol.110.070607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Bhatnagar, V. et al. Analysis of ABCG2 and other urate transporters in uric acid homeostasis in chronic kidney disease: potential role of remote sensing and signaling. Clin Kid J9, 444–453, doi:10.1093/ckj/sfw010 (2016). [DOI] [PMC free article] [PubMed]
  • 49.Liu HC, et al. An Organic Anion Transporter 1 (OAT1)-centered Metabolic Network. J Biol Chem. 2016;291:19474–19486. doi: 10.1074/jbc.M116.745216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Ahn, S. Y. et al. Linkage of Organic Anion Transporter-1 to Metabolic Pathways through Integrated “Omics”-driven Network and Functional Analysis. J Biol Chem 286, 31522–31531 doi:10.1074/jbc.M111.272534 (2011). [DOI] [PMC free article] [PubMed]
  • 51.Ellis RJ, et al. Indoxyl sulphate and kidney disease: Causes, consequences and interventions. Nephrology (Carlton) 2016;21:170–177. doi: 10.1111/nep.12580. [DOI] [PubMed] [Google Scholar]
  • 52.DiNatale BC, et al. Kynurenic acid is a potent endogenous aryl hydrocarbon receptor ligand that synergistically induces interleukin-6 in the presence of inflammatory signaling. Toxicol Sci. 2010;115:89–97. doi: 10.1093/toxsci/kfq024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Mezrich JD, et al. An interaction between kynurenine and the aryl hydrocarbon receptor can generate regulatory T cells. J Immunol. 2010;185:3190–3198. doi: 10.4049/jimmunol.0903670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Schroeder JC, et al. The uremic toxin 3-indoxyl sulfate is a potent endogenous agonist for the human aryl hydrocarbon receptor. Biochemistry-Us. 2010;49:393–400. doi: 10.1021/bi901786x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Jaeger C, Tischkau SA. Role of Aryl Hydrocarbon Receptor in Circadian Clock Disruption and Metabolic Dysfunction. Environ Health Insights. 2016;10:133–141. doi: 10.4137/EHI.S38343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Noakes R. The aryl hydrocarbon receptor: a review of its role in the physiology and pathology of the integument and its relationship to the tryptophan metabolism. Int J Tryptophan Res. 2015;8:7–18. doi: 10.4137/IJTR.S19985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Wang J, et al. Kynurenic acid as a ligand for orphan G protein-coupled receptor GPR35. J Biol Chem. 2006;281:22021–22028. doi: 10.1074/jbc.M603503200. [DOI] [PubMed] [Google Scholar]
  • 58.Khan MZ, He L. Neuro-psychopharmacological perspective of Orphan receptors of Rhodopsin (class A) family of G protein-coupled receptors. Psychopharmacology (Berl) 2017;234:1181–1207. doi: 10.1007/s00213-017-4586-9. [DOI] [PubMed] [Google Scholar]
  • 59.Zhao P, et al. Crucial positively charged residues for ligand activation of the GPR35 receptor. J Biol Chem. 2014;289:3625–3638. doi: 10.1074/jbc.M113.508382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Evans AM, DeHaven CD, Barrett T, Mitchell M, Milgram E. Integrated, nontargeted ultrahigh performance liquid chromatography/electrospray ionization tandem mass spectrometry platform for the identification and relative quantification of the small-molecule complement of biological systems. Anal Chem. 2009;81:6656–6667. doi: 10.1021/ac901536h. [DOI] [PubMed] [Google Scholar]
  • 61.Ohta T, et al. Untargeted metabolomic profiling as an evaluative tool of fenofibrate-induced toxicology in Fischer 344 male rats. Toxicol Pathol. 2009;37:521–535. doi: 10.1177/0192623309336152. [DOI] [PubMed] [Google Scholar]
  • 62.Xia J, Sinelnikov IV, Han B, Wishart DS. MetaboAnalyst 3.0–making metabolomics more meaningful. Nucleic Acids Res. 2015;43:W251–257. doi: 10.1093/nar/gkv380. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental Data (432.7KB, pdf)

Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES